You are on page 1of 21

Cell Metabolism

Review

AMPK and TOR: The Yin and Yang of Cellular


Nutrient Sensing and Growth Control
Asier González,1 Michael N. Hall,1 Sheng-Cai Lin,2 and D. Grahame Hardie3,*
1Biozentrum, University of Basel, CH4056 Basel, Switzerland
2School of Life Sciences, Xiamen University, Xiamen, 361102 Fujian, China
3Division of Cell Signalling & Immunology, School of Life Sciences, University of Dundee, Dundee, DD1 5EH, Scotland, UK

*Correspondence: d.g.hardie@dundee.ac.uk
https://doi.org/10.1016/j.cmet.2020.01.015

The AMPK (AMP-activated protein kinase) and TOR (target-of-rapamycin) pathways are interlinked,
opposing signaling pathways involved in sensing availability of nutrients and energy and regulation of cell
growth. AMPK (Yin, or the ‘‘dark side’’) is switched on by lack of energy or nutrients and inhibits cell growth,
while TOR (Yang, or the ‘‘bright side’’) is switched on by nutrient availability and promotes cell growth. Genes
encoding the AMPK and TOR complexes are found in almost all eukaryotes, suggesting that these pathways
arose very early during eukaryotic evolution. During the development of multicellularity, an additional tier of
cell-extrinsic growth control arose that is mediated by growth factors, but these often act by modulating
nutrient uptake so that AMPK and TOR remain the underlying regulators of cellular growth control. In this re-
view, we discuss the evolution, structure, and regulation of the AMPK and TOR pathways and the complex
mechanisms by which they interact.

All eukaryotic cells are now thought to have arisen via a single need to monitor the output of ATP by mitochondria and to upre-
endosymbiotic event when an archaeal host cell engulfed bacte- gulate their ATP-generating capacity if or when the supply of ATP
ria that were capable of oxidative metabolism, the latter eventu- was insufficient; this is now a major function of the AMPK (AMP-
ally becoming mitochondria (Lane, 2006; Sagan, 1967). This activated protein kinase) signaling pathway. In addition, there
event was followed by the transfer of most of the genes from would have been a requirement to monitor the supply of
the genome of the endosymbiont to that of the host—it has nutrients, such as amino acids and glucose, produced at the
been argued that this separation of energy-generating capacity lysosome by phagocytosis, pinocytosis, or autophagy and to up-
from gene expression allowed a large increase in the energy regulate cell growth when these nutrients were available; this is
available per gene, thus permitting a major expansion in gene now a key function of the TOR (target-of-rapamycin) pathway.
number in the host (Lane and Martin, 2010). This may in turn We propose that these two opposing pathways, which are pre-
have enabled major enhancements in the complexity of eukary- sent in almost all present-day eukaryotes, are the descendants
otic cells compared with their prokaryotic counterparts, of ancient nutrient sensing and signaling pathways that arose
including the development of endomembrane systems such as very early during eukaryotic evolution. AMPK represents the
lysosomes or vacuoles (de Duve, 2005), and the associated traf- Yin (‘‘dark’’ or ‘‘passive’’) side that signals lack of nutrients or
ficking of materials between these internal compartments and insufficient ATP and inhibits cell growth, whereas TOR repre-
the plasma membrane via membrane-bound vesicles. New sents the Yang (‘‘bright’’ or ‘‘active’’) side that signals availability
cellular functions this led to were phagocytosis and pinocytosis, of nutrients and promotes cell growth. Just as in the Chinese phi-
used by many protists today as mechanisms of feeding, and losophy of Taoism from which the Yin-Yang concept is derived,
autophagy, used by all eukaryotic cells for recycling of cellular an appropriate balance between these two opposing elements
components that are damaged or surplus to requirements, or ensures homeostasis and thus a healthy cell or organism.
as an emergency measure during nutrient starvation. Phagocy- In present-day unicellular eukaryotes, including fungi like
tosis, pinocytosis, and autophagy deliver proteins, lipids, and Saccharomyces cerevisiae, growth and proliferation are regu-
carbohydrates, or even whole organelles, such as mitochondria, lated almost entirely by nutrient availability, and the orthologs
to lysosomes or vacuoles; the latter are acidic compartments, of AMPK and TOR play crucial roles in this. However, during
where the engulfed materials are broken down to recycle their the development of multicellular organisms, the uptake (and
components either for catabolism or re-use. Lysosomes or vac- hence the intracellular availability) of nutrients has become
uoles can therefore be considered to be the ‘‘gut’’ or digestive modulated by an additional tier of cell-extrinsic regulation medi-
systems of unicellular eukaryotes, particularly in amoeboid pro- ated by growth factors and cytokines (Palm and Thompson,
tists that feed by phagocytosis or pinocytosis. They would there- 2017). It can be argued that these cell-extrinsic factors ‘‘license’’
fore have been a major source of nutrients and appear to have or allow cells to take up nutrients but that the AMPK and TOR
developed into hubs for nutrient sensing, as discussed below. pathways, which sense intracellular nutrient availability, remain
As these processes were evolving, early eukaryotes would the primary internal regulators of cell growth and proliferation.
have needed signaling pathways that could monitor the function Interestingly, most of the mutations that cause cancer in multi-
of their new internal organelles and regulate cell growth and pro- cellular organisms appear to affect the higher-level, cell-extrinsic
liferation accordingly. For example, there would have been a regulation of cell growth. Such mutations allow cancer cells to

472 Cell Metabolism 31, March 3, 2020 ª 2020 Elsevier Inc.


Cell Metabolism

Review

Table 1. Gene and Protein Names for AMPK, TORC1, and TORC2 Subunit Orthologs in Humans and Different Model Organisms
(Arabidopsis thaliana, Drosophila melanogaster, Caenorhabditis elegans, Saccharomyces cerevisiae, Schizosaccharomyces pombe)
Signaling H. sapiens A. thaliana D. melanogaster C. elegans S. cerevisiae S. pombe
complex Gene Protein Locus Protein Gene Protein Gene Protein Gene Protein Gene Protein
AMPK PRKAA1 a1 AT3G01090 KIN10 snfA AMPKa aak-1 AAK-1 SNF1 Snf1 ssp2 Ssp2
PRKAA2 a2 AT3G29160 KIN11 aak-2 AAK-2 ppk9 Ppk9
PRKAB1 b1 AT5G21170 KINb1 alc AMPKb aakb-1 AAKB-1 SIP1 Sip1 amk2 Amk2
PRKAB2 b2 AT4G16360 KINb2 aakb-2 AAKB-2 SIP2 Sip2
AT2G28060 KINb3a GAL83 Gal83
PRKAG1 g1 AT3G48530 KINgb SNF4Ag AMPKg aakg-1 AAKG-1 SNF4 Snf4 cbs2 Cbs2
PRKAG2 g2 AT1G09020 KINbgc aakg-2 AAKG-2
PRKAG3 g3 aakg-3 AAKG-3
aakg-4 AAKG-4
aakg-5 AAKG-5
TORC1 MTOR mTOR AT1G50030 TOR tor TOR let-363 TOR TOR1/TOR2 TOR1/TOR2 tor2 Tor2
MLST8 mLST8 AT3G18140 LST8-1 lst8 LST8 mlst-8 LST-8 LST8 Lst8 pop3 Pop3
AT2G22040 LST8-2d
RPTOR RAPTOR AT5G01770 RAPTOR1A raptor RAPTOR daf-15 DAF-15 KOG1 Kog1 mip1 Mip1
AT3G08850 RAPTOR1B
TORC2 MTOR mTOR AT1G50030 TOR tor TOR let-363 TOR TOR2 TOR2 tor1 Tor1
MLST8 mLST8 AT3G18140 LST8-1 lst8 LST8 mlst-8 LST-8 LST8 Lst8 pop3 Pop3
AT2G2204 LST8-2d
RICTOR RICTOR - - rictor RICTOR rict-1 RICTOR AVO3 Avo3 ste20 Ste20
MSIN1 mSIN1 - - sin1 SIN1 sinh-1 SIN1 AVO1 Avo1 sin1 Sin1
a
This is an unusual plant-specific b subunit that contains the C-terminal domain but lacks a CBM (carbohydrate-binding module).
b
By sequence, this appears to be an ortholog of mammalian g subunits, but it does not appear to form functional heterotrimers (Zhao, 2019).
c
This is an unusual plant-specific g subunit that contains a CBM (carbohydrate-binding module) fused to the four CBS motifs found in other AMPK-g
subunits.
d
This is probably a non-functional protein (Moreau et al., 2012).

become ‘‘rebels’’ that have partially reverted to their unicellular 3 atypical protein kinases (compared with >500 in humans)
origins and that switch over to using cell-intrinsic growth control, (Miranda-Saavedra et al., 2007). Ancestors of these organisms
based on nutrient availability and controlled by the AMPK and most likely did have AMPK genes, but the modern-day descen-
TOR pathways. dants may have been able to dispense with them because the
host cell would provide AMPK that regulates cellular energy bal-
Yin: The Structure and Regulation of AMPK/SNF1 ance on their behalf. Consistent with this, species closely related
Complexes to P. falciparum that cause malaria in birds (P. gallinaceum and
Subunit Structure and Evolution P. relictum) do still have conventional AMPK genes (Böhme
AMPK appears to occur universally as heterotrimeric complexes et al., 2018). Interestingly, TOR genes are missing in E. cuniculi
comprising catalytic a ubunits and regulatory b and g subunits and P. falciparum (Figure 1) but are also absent in
(Ross et al., 2016b). Genes encoding all three subunits are P. gallinaceum and P. relictum.
readily found within the genomes of almost all eukaryotes Mammals, including humans, have two genes encoding iso-
(Table 1 and Figure 1). However, the orthologs in budding yeast forms of AMPK-a (a1 and a2), two encoding AMPK-b (b1 and
(S. cerevisiae) and plants are not allosterically activated by AMP b2), and three encoding AMPK-g (g1, g2, and g3) (Table 1).
and were discovered independently of mammalian AMPK by ge- These multiple isoforms appear to have arisen during the
netic approaches (Alderson et al., 1991; Celenza and Carlson, two rounds of whole-genome duplication that occurred during
1986). They are therefore not usually referred to as AMPK but the early evolution of vertebrates (Ross et al., 2016b). All
instead in yeast as Snf1 complexes (SNF1 being the gene encod- twelve combinations of these subunit isoforms are able to
ing the catalytic subunit) and in plants as Snf1-related kinase-1 form heterotrimeric complexes, although it is not certain that
(SnRK1) complexes. all combinations exist in vivo. Structures for several almost-
Interestingly, the only eukaryotes known to lack AMPK subunit complete human AMPK heterotrimers, i.e., a2b1g1 (Xiao
orthologs are parasites that spend all or most of their life cycle et al., 2013), a1b1g1 (Calabrese et al., 2014), a1b2g1 (Li
living inside other eukaryotic cells, including Encephalitozoon et al., 2015), and a2b2g1 (Ngoei et al., 2018), have been ob-
cuniculi and Plasmodium falciparum, the latter being the causa- tained via X-ray crystallography. The complexes were all crys-
tive agent of human malaria (Figure 1). These parasitic eukary- tallized in active conformations, and their structures are
otes appear to have undergone stringent selection for small very similar; a schematic representation of a generalized
genome size, with E. cuniculi having one of the smallest known AMPK heterotrimer based on these structures is shown in
genomes of any eukaryote, encoding only 29 conventional and Figure 2.

Cell Metabolism 31, March 3, 2020 473


Cell Metabolism

Review
Figure 1. Conservation of TOR and AMPK
Signaling Components among Eukaryotic
Species
Black boxes indicate presence, and white boxes
absence, of the indicated genes/proteins in the
corresponding organisms (Tatebe and Shiozaki,
2017; van Dam et al., 2011). Gray boxes indicate
limited similarity to the human counterpart. There is
no evidence that the S. cerevisiae Rheb regu-
lates TORC1.

in its promotion of catabolic processes, in-


hibition of anabolic processes, and effects
on DNA replication, are shown in Figure 3.
In the canonical mechanism that is en-
shrined in its name, AMPK is activated by
binding of 50 -AMP, with activation occur-
ring not by one but three mechanisms: (1)
allosteric activation of AMPK already
phosphorylated on Thr172 (Carling et al.,
1987; Ferrer et al., 1985; Yeh et al., 1980);
(2) enhanced Thr172 phosphorylation by
the LKB1 complex (Hawley et al., 1995);
and (3) protection against Thr172 dephos-
Structure of AMPK and Canonical Adenine Nucleotide phorylation by protein phosphatases (Davies et al., 1995). All
(Energy)-Sensing Mechanism three effects are due to binding of AMP to AMPK, not to the up-
Although the main theme of this review is nutrient sensing, we will stream kinase or phosphatase, and this tripartite mechanism en-
first discuss the classical or ‘‘canonical’’ mechanism by which sures that the system responds to small increases in AMP in a
AMPK responds to the changing energy status of cells. The cata- very sensitive manner. Although there is general agreement
lytic a subunits of AMPK contain, at their N-termini, conventional that only AMP binding causes effect #1 above, ADP binding simi-
serine/threonine kinase domains with a small N-lobe and larger C- larly triggers effects #2 and #3 (Oakhill et al., 2011; Xiao et al.,
lobe and the catalytic site in the cleft between them. As with many 2011). However, most AMPK complexes (apart from those con-
other members of the ePK (eukaryotic protein kinase) family, taining the g2 isoform) are about 10-fold more sensitive to AMP
AMPK complexes are only significantly active when phosphory- than ADP, suggesting that increases in AMP are the primary acti-
lated at a critical residue within the activation loop, a stretch of vating signal, although increases in ADP may contribute (Ross
z20 amino acids in the C-lobe between the highly conserved et al., 2016a). All of the activating effects of AMP and ADP are
DFG and APE motifs. In AMPK, the critical phosphorylation site antagonized by binding of ATP so that the AMPK system effec-
is a threonine, usually referred to as Thr172 after its position in tively monitors cellular AMP:ATP and ADP:ATP ratios.
the rat a2 sequence where originally mapped (Hawley et al., Where are the regulatory binding sites where these adenine
1996). Thr172 is not phosphorylated by AMPK itself but by up- nucleotides are sensed? The g subunits contain four tandem re-
stream kinases, principally by LKB1 (liver kinase B1) (Hawley peats of a sequence termed a CBS (cystathionine b-synthase)
et al., 2003; Shaw et al., 2004; Woods et al., 2003), the active motif (Bateman, 1997). These occur, usually as just two tandem
form of which is a heterotrimeric complex also containing repeats, in about 75 proteins in humans, and they are also found
STRAD-a or –b, and the scaffold protein MO25-a or –b (Zeqiraj in archaea and bacteria. Single pairs of tandem CBS repeats
et al., 2009). LKB1 was originally identified as the product of the associate into pseudodimers (termed Bateman modules), poten-
tumor suppressor gene STK11, which is mutated in Peutz-Jegh- tially creating two pseudo-symmetrical ligand-binding sites in
ers Syndrome (an inherited susceptibility to cancer) as well as in the intervening cleft, although in many cases, only one is utilized.
some sporadic (i.e., non-inherited) cancers, especially lung ade- These sites usually bind ligands containing adenosine or (less
nocarcinomas (Alessi et al., 2006; Ji et al., 2007; Sanchez-Ces- often) guanosine (Anashkin et al., 2017; Scott et al., 2004). The
pedes et al., 2002). Although LKB1 was subsequently shown to two Bateman modules in each AMPK-g subunit associate
phosphorylate and activate twelve other kinases with kinase do- head-to-head to form a flattened disk with four potential binding
mains related to AMPK (the AMPK-related kinase family; Jaleel sites for adenine nucleotides in the center (Figure 2). However,
et al., 2005; Lizcano et al., 2004), AMPK was the first downstream only three are utilized, i.e., CBS3, which is accessible from one
target for LKB1 to be identified, and this introduced an intriguing face of the g subunit, and CBS1 and CBS4, accessible from
connection between AMPK and cancer. Indeed, it is now clear the other. The critical site appears to be CBS3; the a-linker, a
that AMPK can also act as a tumor suppressor, at least in certain flexible region of the a subunit that connects the a-AID
animal models of cancer (Vara-Ciruelos et al., 2019). (a-auto-inhibitory domain) and a-CTD (a-C-terminal domain),
A summary of the canonical and non-canonical mechanisms wraps around the face of the g subunit containing CBS3, con-
that activate AMPK, and selected downstream targets involved tacting its bound AMP (Figure 2). This interaction is not thought

474 Cell Metabolism 31, March 3, 2020


Cell Metabolism

Review
Figure 2. Schematic View of the Structure of
the AMPK Heterotrimer
The diagram is a composite derived from the
structures of the human a2b1g1 (Xiao et al., 2013),
a1b1g1 (Calabrese et al., 2014), and a1b2g1 (Li
et al., 2015) complexes and is an active conforma-
tion with Thr172 phosphorylated and three mole-
cules of AMP bound to the g subunit. The a subunit
is shown in yellow (apart from the ST loop, in red),
the b subunit in lilac, and the g subunit in blue-green.
The a-linker is depicted as a yellow chain connect-
ing the a-AID and the a-CTD, and it contacts AMP
bound in the CBS3 site. The ST loop is not resolved
in any of the structures, and its exact positioning is
speculative. The N-terminal regions of the b sub-
units, and the linker between the b-CBM and the
b-CTD, (not shown) are either absent or are not
resolved in any of the structures.

residues equivalent to Thr172 were re-


ported to be inhibited by ADP in
S. cerevisiae (Mayer et al., 2011) and by
AMP in plants (Sugden et al., 1999a). Allo-
steric activation by AMP has been reported,
although not well studied, using the com-
plexes from D. melanogaster (Pan and Har-
die, 2002), C. elegans (Apfeld et al., 2004),
to occur when ATP is bound at CBS3 instead of AMP, and the and S. pombe (Forte et al., 2019). It seems possible that allosteric
consequent release of the a-linker from the g subunit is pro- activation, which is physiologically significant in intact cells
posed to allow the a-AID to rotate back into its inhibitory position (Gowans et al., 2013), was a later evolutionary refinement that
behind the kinase domain (Chen et al., 2009; Chen et al., 2013; Li increased the overall sensitivity of the system to small changes
et al., 2015; Xiao et al., 2011; Xin et al., 2013); this model thus ex- in AMP.
plains allosteric activation by AMP as well as its antagonism by Non-canonical Activation of AMPK by Ligands Binding at
ATP. At the same time, the resulting conformational changes the ADaM Site
may alter the accessibility of Thr172 for phosphorylation and/or The heterotrimeric AMPK complex contains other ligand-binding
dephosphorylation, although those aspects of the mechanism sites whose physiological function remains less clear. One is the
are less well understood. The functions of the CBS1 and CBS4 glycogen-binding site on the b-CBM (b-carbohydrate-binding
sites are less clear, although they are close to the CBS3 site in module), which is present in the b subunits of all eukaryotes
the center of the CBS repeats, where the three sites interact. and in mammalian cells and causes a proportion of AMPK to
One proposal is that CBS1 binds ATP permanently, while bind to glycogen (Hudson et al., 2003; Polekhina et al., 2003;
CBS4 binds AMP permanently, and that these constitutive bind- Polekhina et al., 2005). Intriguingly, as well as a conventional
ing events alter the conformation of the CBS3 site such that it has CBM on the b subunit, many higher plant SnRK1 complexes
a higher affinity for AMP than ADP or ATP (Gu et al., 2017b). This also contain a second CBM fused at the N terminus of the g sub-
helps to explain how AMPK achieves the difficult task of sensing unit, forming a so-called bg subunit (Lumbreras et al., 2001;
changes in AMP in the 30–300 mM range despite the presence of Zhao, 2019). Although it has been proposed that the single
mM concentrations of ATP (Gowans et al., 2013). An additional CBM of mammalian AMPKs may allow them to sense the struc-
explanation is that only Mg2+-free ATP competes with AMP at tural state of glycogen (McBride et al., 2009), more work is
the CBS3 site (Pelosse et al., 2019), although 90% of intracellular required to confirm that hypothesis. Another ligand-binding
ATP is thought to be present at the Mg.ATP2- complex. Accord- site lies in a cleft (termed the ADaM site) between the other
ing to this model, the ATP and AMP constitutively bound at the face of the CBM (i.e., opposite to the glycogen-binding site)
CBS1 and CBS4 sites, respectively, act essentially as regulatory and the N-lobe of the kinase domain on the a subunit (Figure 2).
co-factors. This explains why a functional CBS4 site is required Several ligands that bind in this site cause a dramatic allosteric
for activation even when overall AMP levels remain at the basal activation of AMPK with, usually, a more modest effect to pro-
level (Zong et al., 2019). mote net Thr172 phosphorylation (Göransson et al., 2007;
Although the sequences of the a, b, and g subunits are well Sanders et al., 2007; Scott et al., 2014; Yan et al., 2019). Howev-
conserved, the regulation by adenine nucleotides of AMPK ortho- er, a curious feature is that with the exception of salicylate (a nat-
logs from eukaryotes other than mammals is much less well stud- ural product of plants, but not animals) (Hawley et al., 2012), all of
ied. As mentioned earlier, neither Snf1 complexes from the compounds currently known to bind there are synthetic mol-
S. cerevisiae (Wilson et al., 1996) nor SnRK1 complexes from ecules that emerged from high-throughput screens searching for
plants (Mackintosh et al., 1992) appears to be allosterically acti- allosteric activators of AMPK (e.g., Cokorinos et al., 2017; Cool
vated by AMP, although the dephosphorylation of the threonine et al., 2006; Myers et al., 2017). This binding site is therefore a

Cell Metabolism 31, March 3, 2020 475


Cell Metabolism

Review

Figure 3. Canonical and Non-canonical Mechanisms of AMPK Activation


Proteins shown in green promote activation of AMPK, while proteins shown in red promote inhibition (aldolase is a positive effector when unoccupied by FBP).
Canonical activation by energy stress requires LKB1, occurs in the cytoplasm, and is triggered by increases in AMP:ATP or ADP:ATP ratios. By contrast, non-
canonical activation by glucose starvation involves translocation of AXIN:LKB1 to the lysosome, where a pool of AMPK myristoylated on the b subunit resides
permanently and can occur in the absence of any changes in adenine nucleotides. Non-canonical activation by Ca2+ ions released from the ER or within the
nucleus, triggered by hormones or DNA damage, respectively, requires CaMKK2 and not LKB1. Note that the localized increase in Ca2+ caused by activation of
TRPV channels is not sufficient to activate CaMKK2. See main text for details.

type of ‘‘orphan receptor,’’ and many researchers in the field sus- (inositol-3,4,5-trisphosphate), such as thrombin acting at prote-
pect that there is a unidentified metabolite occurring in animal ase-activated receptor-1 in endothelial cells (Stahmann et al.,
cells that binds to it, hence the acronym ADaM (allosteric drug 2006), acetylcholine acting at M3 muscarinic receptors in various
and metabolite) site (Langendorf and Kemp, 2015). cell types (Jadeja et al., 2019; Merlin et al., 2010; Thornton et al.,
Non-canonical Activation of AMPK by Ca2+ and by DNA 2008; Xue et al., 2016), and ghrelin acting at GHSR1 receptors in
Damage neurons of the hypothalamus (Yang et al., 2011). AMPK is also
Thr172 can also be phosphorylated by alternate upstream ki- activated via a Ca2+/CaMKK2-dependent mechanism by the
nases, including the Ca2+/calmodulin-dependent kinase, growth factor VEGF (vascular endothelial growth factor) acting
CaMKK2 (Hawley et al., 2005; Hurley et al., 2005; Woods et al., at the tyrosine kinase-linked VEGF receptor in endothelial cells,
2005) and TAK1 (Transforming growth factor-b-Activated Ki- which triggers release of IP3 via activation of PLCg (phospholi-
nase-1) (Momcilovic et al., 2006). The physiological importance pase C-g) (Reihill et al., 2007; Stahmann et al., 2010).
of TAK1 as a means of AMPK activation is not well established, Another non-canonical AMPK activation mechanism occurs in
although there is one report that it is involved in AMPK activation response to DNA damage and/or replicative stress (Figure 3),
in response to TRAIL (tumor necrosis factor-related apoptosis- which can be induced by etoposide, hydroxyurea, aphidicolin,
inducing ligand) (Herrero-Martı́n et al., 2009). By contrast, there or ionizing radiation (Fu et al., 2008; Li et al., 2019b; Sanli
is good evidence that AMPK can be activated by the CaMKK2 et al., 2010). Interestingly, the effects of etoposide, hydroxyurea,
pathway in response to hormones or growth factors that trigger or aphidicolin require CaMKK2 but not LKB1, correlate with in-
release of Ca2+ from the endoplasmic reticulum (Figure 3). This creases in nuclear Ca2+, only activate AMPK in the nucleus,
includes hormones acting at G protein-coupled receptors linked and (at least for etoposide) only activate the a1 isoform (Li
via Gq/G11 to release of the Ca2+-mobilizing messenger IP3 et al., 2019b; Vara-Ciruelos et al., 2018). Studies with AMPK

476 Cell Metabolism 31, March 3, 2020


Cell Metabolism

Review

knockout cells reveal that they are hypersensitive to cell death embryo fibroblasts (MEFs) only express AXIN1, AXIN2 is also ex-
induced by DNA damage or replicative stress (Vara-Ciruelos pressed in HEK293T cells so that if AXIN1 expression is knocked
et al., 2018), and this correlates with increased resection of repli- out in HEK293T cells, the lysosomal AMPK activation pathway
cation forks as well as other chromosomal abnormalities (Li et al., remains intact (Zong et al., 2019). In addition, in some cell types
2019b). The defects in the knockout cells have been attributed, that rely on glycolysis for ATP production, glucose starvation
at least in part, to lack of phosphorylation by AMPK of the 50 -30 may also activate AMPK by the canonical AMP-dependent
exonuclease EXO1, which normally causes its association with pathway, rendering the lysosomal activation pathway redun-
14-3-3 proteins, thus restraining its ability to resect replication dant. For example, in HEK293 cells (unlike in MEFs), there are
forks (Li et al., 2019b). Since many of these genotoxic treatments rapid increases in cellular AMP:ATP and ADP:ATP ratios after
are used in cancer therapy, it seems likely that they would be glucose removal even when an alternative carbon source like
more efficacious if administered together with an AMPK inhibi- glutamine is provided (Zhang et al., 2017). In these cells, the ca-
tor, thus preventing the protective effects of AMPK against cell nonical AMP-dependent pathway for AMPK activation operates
death induced by DNA damage or replicative stress. independently of the lysosomal AMP-independent pathway in
Non-canonical Activation of AMPK by Glucose response to glucose starvation (Zong et al., 2019). Thus, studies
Starvation of the lysosomal pathway in some cell types or tissues need to
Recent studies in mammalian cells have revealed, perhaps sur- take into account the possibility not only of expression of
prisingly, that activation of AMPK in response to glucose starvation AXIN1 or AXIN2 but also of changing AMP levels.
can occur via a non-canonical, AMP-independent mechanism. Although the results of Zhang et al. (2014) demonstrated that
The first clues came from administration of siRNAs targeting glucose starvation activated AMPK via the lysosomal pathway
AXIN1 into the tail vein of mice, using adenoviral vectors that direct in mammals, it remained unclear how the presence or absence
expression to the liver. After overnight starvation, animals receiving of glucose was sensed. Pursuing this, it became apparent that
siRNA showed diminished AMPK activation and increased fat stor- aldolase, the glycolytic enzyme that converts FBP (fructose-
age in liver. This led to the discovery that AXIN1, which was initially 1,6-bisphosphate) into triose phosphates, which can also be
identified as a central scaffold protein for Wnt signaling (Zeng et al., associated with the v-ATPase complex, is a direct (physical)
1997), binds constitutively to LKB1 and acts as an adaptor for sensor for FBP. When aldolase is unoccupied by FBP (whose
LKB1 to associate with and phosphorylate AMPK; this initial char- levels rapidly decrease upon glucose deprivation), the v-
acterization of the role of AXIN1 was based on an in vitro reconsti- ATPase complex undergoes conformational changes that
tution experiment where high levels of AMP were required for the inhibit its activity as a proton pump (as suggested by increased
interaction to occur (Zhang et al., 2013), which can now be classi- pH levels in the lysosomal lumen; Zhang et al., 2017) and also
fied as a cytosolic, AXIN/AMP-dependent mechanism (Zong et al., allow the AXIN1:LKB1 complex to interact with the v-ATPase
2019). A subsequent yeast two-hybrid screen searching for novel and Ragulator. Multiple lines of evidence support the idea
AXIN1-interacting proteins (Zhang et al., 2014) identified p18/ that aldolase is the direct sensor. First, knockdown of all iso-
LAMTOR1, a protein anchored to the lysosomal membrane by forms of aldolase caused constitutive activation of AMPK,
N-terminal myristoyl and palmitoyl modifications (Nada et al., even in high glucose. Second, in cells expressing the D34S
2009). p18/LAMTOR1 is a key component of the Ragulator com- mutant of aldolase, which has a greatly reduced kcat despite
plex, which (as will be discussed later) plays a central role in the an almost unchanged Km for FBP (Morris and Tolan, 1993)
activation of mTORC1 via interaction with the vacuolar ATPase (meaning that FBP will accumulate in the active site of aldolase
(v-ATPase) (Bar-Peled et al., 2012; Sancak et al., 2010; Zoncu even in low glucose), AMPK was not activated by glucose star-
et al., 2011). In LAMTOR1 null cells or cells with knockdown of vation (Zhang et al., 2017). Importantly, this mechanism for
the v0c subunit of the v-ATPase, AMPK activation induced by AMPK regulation by glucose can occur in the absence of any
glucose starvation was no longer observed. In addition, AXIN1, changes in adenine nucleotide ratios. For example, in MEFs
in complex with LKB1, was found to translocate to the lysosomal transferred from media with high glucose (25 mM) to medium
surface, forming a supramolecular complex with the Ragulator containing glucose concentrations around 5 mM or in livers of
and v-ATPase, which was not observed in LAMTOR1 null cells or mice starved overnight (when blood glucose dropped from 9
cells with knockdown of the v-ATPase v0c subunit (Zhang et al., to 3 mM), AMPK was activated without any associated
2014). By this mechanism, LKB1 is brought to the vicinity of a changes in cellular AMP:ATP or ADP:ATP ratios. Interestingly,
pool of AMPK that appears to permanently reside on the lysosomal however, if glutamine (the other major carbon source in the me-
membrane due to N-terminal myristoylation of the b subunit. This dium) was removed from the medium as well as glucose, there
overall mechanism is now referred to as the lysosomal AMPK acti- was an additional, delayed (but ultimately larger) activation of
vation pathway (Figure 3). AMPK that did correlate with increases in AMP:ATP and AD-
It should be noted that AXIN has two isoforms, AXIN1 and P:ATP ratios (Zhang et al., 2017). These results indicate that
AXIN2, which are functionally redundant both in Wnt signaling the non-canonical glucose-sensing mechanism for AMPK acti-
(Chia and Costantini, 2005) and in the lysosomal AMPK activa- vation can act in parallel with the canonical AMP-dependent
tion pathway (Zong et al., 2019). While AXIN1 is ubiquitously ex- mechanism. In line with the concept that glucose availability
pressed, AXIN2 is mainly expressed in neuronal cells and some can be sensed independently of cellular energy status, neither
actively proliferating cells. For example, AXIN2 is not expressed pyruvate nor glutamine, which both feed into the TCA cycle for
in differentiated hepatocytes (Zong et al., 2019), except for a ATP production, prevents lack of glucose from activating
small population of self-renewing cells adjacent to the central AMPK. Indeed, it is now clear that the AXIN/lysosome-depen-
vein in the liver lobule (Wang et al., 2015a). Similarly, while mouse dent and AMP-dependent mechanisms can co-exist, with their

Cell Metabolism 31, March 3, 2020 477


Cell Metabolism

Review

contributions to overall AMPK activation depending on the Since activation by starvation for key carbon sources (espe-
magnitude of any increases in AMP as well as the subcellular cially glucose) appears to be a common feature of the AMPK or-
location (Zong et al., 2019). thologs from mammals, plants, and budding yeast, yet they differ
Another recent study has uncovered the mechanism that in their regulation by adenine nucleotides, it is tempting to spec-
signals the presence or absence of FBP in the active site of ulate that sensing of glucose rather than energy may have been
aldolase to the formation of the AXIN-LKB1-AMPK complex on the ancestral role of the kinase. However, it remains unclear
the lysosomal membrane. It was demonstrated that TRPV (tran- exactly how carbon starvation causes activation of the orthologs
sient receptor potential V) channels located on the ER (endo- in plants and yeast.
plasmic reticulum) membrane are required for AMPK activation Downstream Targets of AMPK
in response to low glucose. The current model is that aldolase AMPK phosphorylates downstream targets containing well-
that is unoccupied by FBP interacts with TRPV at lysosome:ER defined recognition motifs, and at least 60 have now been well
contact sites, inhibiting its Ca2+-releasing activity. Once the validated—a full discussion of these is beyond the scope of
Ca2+ concentrations at the ER-lysosome contact sites falls this article, and readers are referred to a previous review (Hardie
below a certain level, TRPV gains affinity for the v-ATPase, re- et al., 2016). In general, AMPK phosphorylates and activates
configuring its association with aldolase and causing the forma- proteins involved in catabolic pathways, thus enhancing ATP
tion of the AXIN-based complex to activate AMPK (Li et al., synthesis, while phosphorylating and inactivating proteins
2019a). It should be pointed out that the concentration of the involved in anabolic (biosynthetic) pathways, thus inhibiting cell
TRPV-released Ca2+ (<1 mM) is well below that required for acti- growth while conserving ATP. AMPK also causes a cell cycle ar-
vation of CaMKK2, which is not involved in the lysosomal AMPK rest in G1 phase (Fogarty et al., 2016; Imamura et al., 2001),
activation mechanism. It has been proposed that the pool of although in that case, the direct downstream targets responsible
Ca2+ at the ER-lysosome contact sites acts as a kind of buffer for the effect are not clear. In this section, we will mention only a
or damper, smoothing the output and thus preventing fluctua- few key targets that are important for the effects of AMPK on
tions in AMPK caused by rapid oscillations of FBP binding in catabolic and anabolic pathways.
the active site of aldolase (Li et al., 2019a). Starting with effects on catabolism, in many cell types, AMPK
Glucose starvation also causes rapid activation of the Snf1 activation increases glucose uptake via effects on the trafficking
complex in S. cerevisiae (Wilson et al., 1996; Woods et al., of the glucose transporters, GLUT1 (Barnes et al., 2002) or
1994), and, intriguingly, complexes containing Sip1 (one of three GLUT4 (Kurth-Kraczek et al., 1999). This is achieved in part via
b subunit orthologs in yeast) translocate to the vacuolar mem- phosphorylation and consequent degradation of TXNIP, an a-ar-
brane upon glucose removal (Vincent et al., 2001). However, restin family member that normally promotes reuptake of GLUT1
the detailed mechanism appears to be different from that in and GLUT4 from the plasma membrane by endocytosis (O’Don-
mammalian cells, because no clear AXIN orthologs are found nell and Schmidt, 2019; Wu et al., 2013). In the case of GLUT4,
in yeast. Once activated, the Snf1 complex phosphorylates the AMPK also phosphorylates TBC1D1, a GTPase activating pro-
transcriptional repressor Mig1 (Smith et al., 1999; Treitel et al., tein (GAP) for members of the Rab family, causing dissociation
1998), triggering both its inactivation (Papamichos-Chronakis of TBC1D1 from intracellular GLUT4-storage vesicles (GSVs)
et al., 2004) and nuclear export (DeVit and Johnston, 1999). with consequent conversion of Rabs to their GTP-bound forms,
Mig1 binds to and inhibits the promoters of many glucose- thus promoting trafficking of GSVs to the plasma membrane
repressed genes, including the SUC2 gene encoding a secreted (Pehmøller et al., 2009). AMPK can also phosphorylate and acti-
invertase that is required to metabolize alternate carbon sour- vate 6-phosphofructo-2-kinase, the enzyme that generates fruc-
ces, such as sucrose or raffinose (Hedbacker and Carlson, tose-2,6-bisphosphate, a potent allosteric activator of the key
2008). As in mammalian cells, the Snf1 complex also phosphor- glycolytic enzyme 6-phosphofructo-1-kinase. However, this ef-
ylates and inactivates acetyl-CoA carboxylase, potentially inhib- fect is cell-type-dependent because only the PFKFB2 (Marsin
iting fatty-acid biosynthesis under glucose-limiting conditions et al., 2000) or PFKFB3 (Marsin et al., 2002) isoforms of 6-phos-
(Mitchelhill et al., 1994; Woods et al., 1994). phofructo-2-kinase, which are not expressed ubiquitously, are
Although the effects of starvation for a carbon source are less direct targets for AMPK. AMPK also acutely promotes fatty-
well studied in plants, knockout or silencing of the genes encod- acid oxidation by phosphorylating and inactivating the mito-
ing the AMPK-a orthologs in the moss Physcomitrella patens chondrial isoform of ACC2 (acetyl-CoA carboxylase-2),
(Thelander et al., 2004) and the higher plant Arabidopsis thaliana thus reducing the local pool of malonyl-CoA, an inhibitor of up-
(Baena-González et al., 2007) causes failure to respond appro- take of fatty acids into mitochondria via the transport system
priately to periods of darkness, the equivalent of starvation in involving carnitine palmitoyl-CoA transferase-1 (Winder and Har-
plants. In cells of A. thaliana, the AMPK-a ortholog KIN10 is die, 1996).
responsible for triggering extensive reprogramming of transcrip- In the longer term, AMPK activation tends to promote the
tion affecting thousands of genes, some of which are required for oxidative metabolism typical of quiescent cells, rather than the
adaptive responses like starch breakdown during starvation rapid glucose uptake and glycolysis typical of cells undergoing
(Baena-González et al., 2007; Baena-González and Sheen, rapid proliferation, including tumor cells. First, it promotes mito-
2008). SnRK1 complexes also phosphorylate and inactivate chondrial biogenesis (Zong et al., 2002) as well as expression of
both sucrose phosphate synthase and HMG (3-hydroxy-3-meth- oxidative enzymes (Winder et al., 2000), perhaps by direct phos-
ylglutaryl)-CoA reductase, potentially inhibiting the anabolic phorylation (Ja €ger et al., 2007) or deacetylation (Cantó et al.,
pathways of sucrose and sterol synthesis (Nukarinen et al., 2009) of the transcriptional co-activator, PGC-1a. Second,
2016; Sugden et al., 1999b). AMPK maintains the cellular content of functional, healthy

478 Cell Metabolism 31, March 3, 2020


Cell Metabolism

Review

mitochondria by promoting both mitophagy, via phosphorylation TOR was originally identified genetically in S. cerevisiae via
of the autophagy kinase ULK1 (Unc-51-like kinase 1) (Egan et al., mutations that render cells resistant to the growth-inhibitory
2011b), and mitochondrial fission, perhaps via phosphorylation properties of the antibiotic rapamycin (Heitman et al., 1991;
of proteins involved in mitochondrial fission like MFF (mitochon- Kunz et al., 1993). It was identified in mammalian cells shortly
drial fission factor) or MTFR1L (mitochondrial fission regulator-1- thereafter (Brown et al., 1994; Chiu et al., 1994; Sabatini et al.,
like) (Ducommun et al., 2015; Schaffer et al., 2015; Toyama et al., 1994; Sabers et al., 1995), and the name mTOR (mammalian
2016). Because mitochondria can exist in cells as elongated TOR) was eventually adopted based on the yeast precedent.
branching networks that can be of lengths close to that of the More recently, the HUGO Gene Nomenclature Committee
cell diameter, mitochondrial fission may be necessary to break changed the definition of the mTOR acronym to ‘‘mechanistic
these networks down into smaller segments suitable for mitoph- TOR’’ in order to create a common nomenclature for TOR in ver-
agy. Consistent with this, the phenotypes of muscle-specific tebrates (Hall, 2013). However, this has led to TOR from nema-
double knockouts of a1/a2 (Lantier et al., 2014) or b1/b2 (O’Neill todes or even yeast sometimes being referred to as mTOR.
et al., 2011) in mice include exercise intolerance associated with TOR forms two structurally and functionally distinct multipro-
the appearance in electron micrographs of mitochondria of tein complexes termed TOR complexes 1 and 2 (TORC1 and
abnormal size and morphology. TORC2), of which only TORC1 is acutely sensitive to rapamycin
Along with these effects on catabolism, AMPK acutely (Loewith et al., 2002). The two TOR complexes, like TOR itself,
switches off most anabolic pathways. It was discovered for its are conserved from yeast to humans, although TORC1 appears
ability to phosphorylate and inactivate ACC1 (acetyl-CoA to be absent from ciliates and TORC2 from plants (Tatebe and
carboxylase-1) and HMG-CoA reductase, two key enzymes of Shiozaki, 2017; van Dam et al., 2011) (Figure 1). In mammals,
fatty acid and cholesterol synthesis, respectively (Hardie et al., mTOR and the adaptor protein mLST8 (mammalian lethal with
1989). Indeed, phosphorylation of ACC1 at Ser80 (Ser79 in ro- SEC13 protein 8) are common to both TOR complexes. RAPTOR
dents), monitored using phosphospecific antibodies, remains (regulatory-associated protein of TOR) is the defining subunit of
the most widely used biomarker for AMPK activation in intact mTORC1, whereas RICTOR (rapamycin-insensitive companion
cells. Moreover, mice with knock-in Ser/Ala mutations of the of mTOR) and mSIN1 (stress-activated MAP kinase interacting
AMPK sites on ACC1 and ACC2 (Fullerton et al., 2013) or protein 1) define mTORC2.
HMG-CoA reductase (Loh et al., 2018) have elevated levels of tri- The domain organization of TOR is also conserved. The C-ter-
glycerides and cholesterol, respectively, demonstrating that minal half of TOR contains a FAT (FRAP, ATM, and TRRAP)
these phosphorylation sites have regulatory significance in vivo. domain followed by the FRB (FKBP-rapamycin binding) domain,
AMPK also switches off glycogen synthesis via phosphorylation the catalytic kinase domain, and a C-terminal FAT domain
of the GYS1 (Jørgensen et al., 2004) and GYS2 (Bultot et al., termed FATC. Structural biologists often refer to the FAT, FRB,
2012) isoforms of glycogen synthase, nucleotide synthesis via kinase, and FATC domains collectively as the FATKIN region
phosphorylation of the PRPS-1 and –2 isoforms of phosphoribo- (Baretic et al., 2016; Imseng et al., 2018) (Figure 4). FATKIN re-
syl pyrophosphate synthetase (Qian et al., 2018), and ribosomal gions are found in all PIKK family members, although only the
RNA synthesis via phosphorylation of TIF-1A/RRN3, a transcrip- FRB domain in TOR binds the FKBP-rapamycin complex. All
tion factor for RNA polymerase-1 (Hoppe et al., 2009). Finally, PIKKs contain long, N-terminal extensions that serve as docking
AMPK switches off the elongation step of protein synthesis in surfaces for binding partners. The N-terminal half of TOR con-
part via phosphorylation of elongation factor-2 kinase (Johanns sists of tandem arrays of HEAT (Huntingtin, elongation factor 3,
et al., 2017), an atypical Ca2+-dependent kinase that phosphor- PP2A, and TOR) and TPR (tetratricopeptide) repeats. The
ylates elongation factor-2 and causes pausing in elongation. HEAT repeats of mTOR bind RAPTOR (Hara et al., 2002; Kim
Other effects on protein synthesis are mediated indirectly by et al., 2002), which also has several characteristic regions: the
inactivation of mTORC1, which is discussed in more detail in a RAPTOR N-terminal conserved (RNC) CASPase-like domain, a
separate section below. central set of seven ⍺-helical repeats termed the armadillo
(ARM) domain, and a C-terminal seven-bladed WD40 b-propel-
Yang – the Structure and Regulation of TOR Complexes ler (Hara et al., 2002; Kim et al., 2002). By contrast, mLST8 is a
Subunit Structure and Evolution small protein consisting entirely of a WD40 b-propeller.
TOR is a serine/threonine protein kinase belonging to the PIKK Structure of the mTORC1 Complex
(phosphatidylinositol kinase-related kinase) family, which also TORC1 architecture was solved by a combination of X-ray
includes DNA-PK and ATM (Keith and Schreiber, 1995). TOR is crystallography and cryo-EM (cryo-electron microscopy) on
conserved in all eukaryotes except (as for AMPK) in the case truncated mTOR-mLST8 (Yang et al., 2013), RAPTOR from the
of a few obligate intracellular parasites like E. cuniculi and fungus Chaetomium thermophilum (Aylett et al., 2016) or the
P. falciparum (Tatebe and Shiozaki, 2017; van Dam et al., plant A. thaliana (Yang et al., 2017), and TOR-Lst8 from the fun-
2011) (Figure 1), which may be able to exploit TOR signaling in gus Kluyveromyces marxianus (Baretic  et al., 2016). These
the host cell. Whereas most eukaryotes contain a single TOR studies described mTORC1 at 4.4  A (Yang et al., 2016) and 3.0
gene, a few possess more than one; for example, budding yeast 
A resolution (Yang et al., 2017), mTORC1 in complex with
(S. cerevisiae) and fission yeast (S. pombe) have two (Shertz FKBP-rapamycin at 5.9  A (Aylett et al., 2016), and mTORC1
et al., 2010) (Table 1), while trypanosomes have up to four bound to its activator RHEB at 3.4  A (Yang et al., 2017).
(Saldivia et al., 2013). Early eukaryotes presumably possessed mTORC1 is a 1 MDa homodimer of heterotrimers (each of the
a single TOR gene that was duplicated and/or lost multiple times latter containing mTOR, RAPTOR, and mLST8) that adopts a
during evolution (Shertz et al., 2010). rhomboidal (lozenge) shape with a large central cavity (Figure 4).

Cell Metabolism 31, March 3, 2020 479


Cell Metabolism

Review

Figure 4. Human mTORC1 Architecture


(A) Linear representation of the domain organization of mTOR, RAPTOR, and mLST8. The residue numbers indicate the domain boundaries. Grey areas in
RAPTOR indicate regions presumed to be disordered linkers, comprising amino acids 687–805 and 841–949.
(B) Cryo-EM derived model of human mTORC1 (PDB: 6BCX) (Yang et al., 2017), with domains colored according to the primary structure scheme in (A). Key
residues for mTORC1 activation at the catalytic site (Asp2338, His2340, Asn2343, and Asp2357; Yang et al., 2013) are highlighted in red, while the two copies of
the TOS peptide of 4EBP are shown in purple. A gray line indicates the RAG binding region. Gray dashed lines represent the two disordered linker regions in
RAPTOR. AMPK, PKA, and NLK phosphorylate RAPTOR at Ser722 plus Ser792, Ser791, and Ser863, respectively. EP300 acetylates RAPTOR at Lys1097
(residue highlighted in magenta). RHEB binds the N terminus and FAT domain of mTOR, distal to the catalytic site (not shown). See main text for details.

It exhibits two-fold (C2) symmetry with the axis of symmetry TOR and (2) TOR from K. marxianus (Baretic  et al., 2016) and hu-
passing through the central cavity. The FATKIN region of each mans (Aylett et al., 2016) are architecturally identical. The horn
of the two copies of mTOR forms a compact unit located near and bridge, in addition to forming the dimer interface, are
the central cavity, on opposite sides of the C2 axis. The two FAT- exposed, suggesting an additional role in binding regulatory or
KIN regions come close to each other but make little or no con- accessory proteins. mLST8 binds to the kinase domain of
tact. Each kinase site is located at the bottom of a deep catalytic mTOR and thereby constitutes the ends of the short axis of the
cleft that is partly obscured by surrounding structural elements, mTORC1 rhomboid. RAPTOR has an extended Z-like shape
suggesting that the kinase activity is regulated by physically re- with the RNC domain and WD40 b-propeller located at opposite
stricting access to the catalytic site (Yang et al., 2017; Yang ends, connected by the ARM domain (Aylett et al., 2016; Yang
et al., 2013). The HEAT repeats of each mTOR subunit form et al., 2017). RAPTOR also contributes to the mTORC1 dimer
two distinct helical solenoids, one a low-curvature bridge/M- interface, because the ARM domain of one RAPTOR binds the
HEAT (hereafter referred to as the ‘‘bridge’’) and the second a horn of one mTOR molecule and the bridge of the other, thereby
high-curvature horn/spiral/N-HEAT (hereafter referred to as the linking the two copies of mTOR. The RAPTOR b-propeller do-
‘‘horn’’) peripherally linked to the bridge (Aylett et al., 2016; Bare- mains are at the ends of the long axis of mTORC1.
 et al., 2016; Yang et al., 2017). The horn of one copy of mTOR
tic Importantly, RAPTOR is also required for mTORC1 substrate
packs against the bridge of the other to mediate dimerization and recruitment. The region in RAPTOR responsible for substrate
form the central cavity. The two-fold symmetry is likely binding is in a cleft between the RNC and the ARM domains,
conserved among TORC orthologs because (1) there is a high located z65  A from the catalytic site (Figure 4) (Yang et al.,
degree of conservation throughout the HEAT repeat region of 2017), via which RAPTOR binds a sequence of five amino acids

480 Cell Metabolism 31, March 3, 2020


Cell Metabolism

Review

termed the TOS (TOR signaling) motif. The TOS motif is defined plex (Kogan et al., 2010; Levine et al., 2013; Powis et al., 2015;
as FXV[E/D]V, where V is a hydrophobic residue and X any res- Zhang et al., 2012). Clearly, the lysosome or vacuole is the
idue (Gouw et al., 2018; Nojima et al., 2003; Schalm and Blenis, TORC1 signaling hub in all eukaryotic cells. Amino-acid suffi-
2002; Yang et al., 2017). TOS motifs are present in some TORC1 ciency promotes the TORC1-activating conformation of the
substrates, such as ribosomal protein S6 kinase (S6K; TOS motif RAG-Gtr heterodimer (RAGA/B or Gtr1 loaded with GTP, and
FDIDL) and eukaryotic translation initiation factor 4E binding pro- RAGC/D or Gtr2 loaded with GDP). In mammals, the active
tein (4EBP; TOS motif FEMDI) (Nojima et al., 2003; Schalm and RAG heterodimer binds RAPTOR and thereby recruits mTORC1
Blenis, 2002; Schalm et al., 2003). However, the mTORC1 sub- from the cytosol to the lysosomal surface, while in budding
strates ULK1 (Dunlop and Tee, 2013) and TFEB (transcription yeast, TORC1 is constitutively bound to the vacuolar surface,
factor EB) (Roczniak-Ferguson et al., 2012; Settembre et al., and the active Gtr1-Gtr2 heterodimer binds Kog1 (yeast ortholog
2012) interact with RAPTOR yet lack an obvious TOS motif. of RAPTOR) to stimulate TORC1 via an unknown mechanism
Furthermore, although the TOS-binding region of RAPTOR is (Binda et al., 2009). From yeast two-hybrid experiments, it has
highly conserved from yeast to mammals, TORC1 substrates been proposed that a region of Kog1 comprising amino acids
in lower eukaryotes seem to lack TOS motifs, so that it is unclear 777–814 in the central ARM domain interacts with Gtr1 (Sekigu-
how TORC1 recognizes its substrates in those organisms. chi et al., 2014). The region in Kog1 is conserved in RAPTOR
Inhibition of TOR by rapamycin depends on the formation of a (amino acids 777–814 in Kog1 correspond to amino acids 595–
complex between rapamycin and the cytoplasmic immunophilin 632 in RAPTOR). Consistent with this, recent structural analyses
FKBP12 (FK506-binding protein of 12 KDa) (Benjamin et al., of RAGAGTP-RAGCGDP in complex with mTORC1 (Anandapada-
2011). An FKBP-rapamycin complex binds the FRB domain at manaban et al., 2019) or with RAPTOR-Ragulator (Rogala et al.,
the lip of the TOR catalytic cleft, forming a lid that physically pre- 2019) revealed that the region in RAPTOR comprising amino
vents access of substrates to the catalytic site. FKBP-rapamycin acids 546–650 binds RAGAGTP. Two additional regions of
does not induce a conformational change in mTOR, suggesting RAPTOR, located between the ARM and WD40 b-propeller do-
that FKBP-rapamycin indeed acts by obstructing substrate ac- mains, interact with RAGCGDP (Rogala et al., 2019). One region
cess (Aylett et al., 2016; Yang et al., 2017; Yang et al., 2013). comprises amino acids 795–806 and the other, amino acids
TORC2 is not acutely inhibited by rapamycin, because the 916–937. The last has been referred to as the ‘‘RAPTOR claw’’
FKBP-rapamycin binding site in the TOR FRB domain in due to its shape (Rogala et al., 2019). Interestingly, it has been
TORC2 is masked by RICTOR (Chen et al., 2018; Gaubitz suggested that the stress-activated MAP kinase-related kinase
et al., 2015; Karuppasamy et al., 2017; Stuttfeld et al., 2018). NLK (Nemo-like kinase) phosphorylates RAPTOR at Ser863,
Cryo-EM studies have resolved S. cerevisiae (Karuppasamy thereby disrupting RAG-RAPTOR interaction and inhibiting
et al., 2017) and human (Chen et al., 2018; Stuttfeld et al., mTORC1 (Yuan et al., 2015). Ser863 is in a structurally unsolved
2018) TORC2 at intermediate resolution. The two mTOR com- and thus presumably disordered linker region (residues 841 to
plexes share many features, including C2 symmetry, similar 949) between the ARM and WD40 b-propeller domains that con-
binding sites for RAPTOR and RICTOR, and a deep catalytic tains several phosphorylation sites (Foster et al., 2010; Wang
cleft. However, full structural interpretation of mTORC2 awaits et al., 2009) (Figure 4) (see below).
higher resolution structural data. The nucleotide binding status of the RAGs is tightly regulated
Regulation of mTORC1 by Lysosomal Recruitment and by conserved GAPs (GTPase activator proteins) and GEFs (gua-
Growth Factors nine nucleotide exchange factors) (González and Hall, 2017; Ni-
TOR controls cell growth and metabolism in response to nutri- castro et al., 2017; Wolfson and Sabatini, 2017) (Figure 5). The
ents, growth factors, and (in part through AMPK) cellular energy heterotrimeric GATOR1 (GAP activity toward RAGs-1) complex
status. Nutrients, especially amino acids, are likely to be the is the GAP for RAGA/B and thus negatively regulates mTORC1
ancestral TORC1 activating inputs, as they are sufficient to acti- activity (Bar-Peled et al., 2013; Panchaud et al., 2013a; Shen
vate TORC1 in unicellular organisms such as yeast. However, in et al., 2018; Shen et al., 2019). GATOR1 is tethered to the lyso-
multicellular organisms, TORC1 activation requires additional somal surface by KICSTOR (KPTN, ITFG2, C12orf66, and
input from growth factors. Mechanistically, amino acid and SZT2-containing regulator of mTORC1) (Peng et al., 2017; Wolf-
growth factor inputs converge on mTORC1 as follows: (1) amino son et al., 2017). The heteropentameric GATOR2 complex acti-
acids stimulate translocation of mTORC1 from the cytosol to the vates mTORC1 by binding and negatively regulating GATOR1
lysosome, where they encounter the small G protein RHEB (RAS via an undefined mechanism (Bar-Peled et al., 2013; Panchaud
homolog enriched in brain), and (2) growth factors activate lyso- et al., 2013b). The lysosomal amino-acid transporter SLC38A9
somal RHEB, enabling it to activate mTORC1 in turn (see below). (Jung et al., 2015; Rebsamen et al., 2015; Wang et al., 2015b;
Amino acid availability is transmitted to TORC1 mainly via the Wyant et al., 2017) acts as a GEF for RAGA (Shen and Sabatini,
RAGs (Ras-related family of small GTPases) (González and Hall, 2018). The Ragulator complex, which was initially described as
2017; Nicastro et al., 2017; Wolfson and Sabatini, 2017) the GEF for RAGA/B (Bar-Peled et al., 2012), is now proposed
(Figure 5). There are four RAGs in mammals (RAGA through instead to activate mTORC1 by accelerating the release of
RAGD) and two in S. cerevisiae (Gtr1 and Gtr2) that form obligate GTP from RAGC (Shen and Sabatini, 2018), while the identity
heterodimers of RAGA or RAGB with RAGC or RAGD and Gtr1 of the GEF for RAGC/D remains unclear. FLCN (folliculin),
with Gtr2. RAGs are attached to the lysosome in mammalian together with its binding partners FNIP1 and FNIP2 (folliculin-in-
cells through the pentameric Ragulator complex (Bar-Peled teracting protein 1 and 2), has been identified as the GAP for
et al., 2012; Sancak et al., 2010), while the Gtr1-Gtr2 heterodimer RAGC/D and thus positively regulates mTORC1 (Petit et al.,
is attached to the vacuole in yeast through the trimeric Ego com- 2013; Tsun et al., 2013).

Cell Metabolism 31, March 3, 2020 481


Cell Metabolism

Review

Figure 5. Cross-Talk between mTORC1 and AMPK Signaling Pathways in Mammals


Proteins shown in green promote activation of mTORC1 (blue box), while proteins shown in red promote its inhibition. Inputs into mTORC1 from AMPK signaling
are shown in gray, because AMPK and mTORC1 would not be simultaneously active. Dashed lines indicate indirect interactions. Amino acids and growth factors
activate mTORC1, which then promotes cell growth by stimulating anabolic processes. Growth-factor-stimulated PI3K activates mTORC2 (yellow box) by
promoting its association with the ribosome. Active mTORC2 then promotes cell proliferation and survival. See main text for details.

Upon amino-acid starvation, the RAG heterodimer assumes TORC1 signaling pathway (Ben-Sahra and Manning, 2017; Guri
an inactive configuration (RAGA/B loaded with GDP and and Hall, 2016; Kim and Guan, 2019). Growth factors like insulin
RAGC/D with GTP) that is unable to recruit mTORC1 to the lyso- bind to RTKs (receptor tyrosine kinases) to activate PI3K (phos-
somal surface so that mTORC1 remains cytosolic and inactive. It phatidylinositol-4,5-bisphosphate 3-kinase), thereby gener-
has been proposed also that the ‘‘inactive’’ conformation of the ating PIP3 (phosphoinositide 3, 4, 5-trisphosphate) (Figure 5).
RAG heterodimer recruits TSC2 (tuberous sclerosis complex 2) PIP3 then co-recruits PDK1 (phosphoinositide-dependent ki-
to the lysosome to inhibit mTORC1 (Demetriades et al., 2014; nase-1) and AKT via their PIP3-binding PH (pleckstrin homol-
Demetriades et al., 2016; Menon et al., 2014). In budding yeast, ogy) domains to the plasma membrane, where PDK1 phos-
glucose withdrawal triggers a Gtr-dependent formation of a vac- phorylates Thr308 in the activation loop of AKT. Activated
uole-associated cylindrical filament of TORC1 molecules, AKT in turn phosphorylates TSC2 on multiple sites to induce
termed a TOROID (TORC1 organized in inhibited domains). the release of the heterotrimeric TSC complex from the lyso-
TOROID formation leads to TORC1 inactivation, and low-resolu- some (Inoki et al., 2002; Menon et al., 2014). The TSC complex
tion cryo-EM reconstructions suggest that this oligomerization consists of TSC1, TSC2, and TBC1D7, and it acts as a GAP to-
causes steric occlusion of the TORC1 active site (Prouteau ward RHEB (Dibble et al., 2012). Reduced TSC complex GAP
et al., 2017). It is not known whether mTORC1 forms TOROID- activity at the lysosome leads to an increase in activated,
like structures. GTP-loaded RHEB, which then binds the N terminus and FAT
As discussed in the introduction to this review, it is thought domain of mTOR to allosterically realign residues in the cata-
that growth factor signaling co-evolved with multicellularity, at lytic site and activate mTORC1 (Chao and Avruch, 2019; Long
which time it was grafted onto the ancestral nutrient-activated et al., 2005; Yang et al., 2017).

482 Cell Metabolism 31, March 3, 2020


Cell Metabolism

Review

Amino-Acid Sensors and pyrimidine synthesis by phosphorylating and activating


Leucine, arginine, and glutamine are among the most effective CAD (carbamoyl-phosphate synthetase 2, aspartate transcarba-
amino acids for activation of mTORC1 (Figure 5). The identity mylase, and dihydroorotase) via S6K (Ben-Sahra et al., 2013; Ro-
of the amino acid sensors upstream of TORC1 has begun to bitaille et al., 2013). Furthermore, mTORC1 promotes lipogenic
emerge recently (Wolfson and Sabatini, 2017). The cytoplasmic gene expression by activating the SREBP (sterol-regulatory
proteins SESTRIN2 (Chantranupong et al., 2014; Kim et al., element-binding protein) transcription factor (Ben-Sahra and
2015; Parmigiani et al., 2014; Saxton et al., 2016b; Saxton Manning, 2017). mTORC1 also inhibits autophagy by phosphory-
et al., 2016c; Wolfson et al., 2016) and CASTOR (cellular arginine lating the autophagy-inducing kinase ULK1 (Kim et al., 2011) and
sensor for mTORC1) (Chantranupong et al., 2016; Saxton et al., TFEB (transcription factor EB) (Martina et al., 2012; Roczniak-Fer-
2016a; Xia et al., 2016) bind and transmit the availability of guson et al., 2012; Settembre et al., 2012). Phosphorylated TFEB
leucine and arginine, respectively, to mTORC1 via the GATOR remains cytosolic and inactive, thus failing to induce expression of
complexes. Under conditions of leucine and arginine depriva- genes required for autophagy and lysosome biogenesis (Puertol-
tion, SESTRIN2 and CASTOR1 bind and most likely inhibit lano et al., 2018) (Figure 5).
GATOR2 upstream of mTORC1. However, growth-promoting S6K has several substrates, including ribosomal protein S6
levels of leucine and arginine disrupt the interactions of SES- and insulin receptor substrate 1 (IRS1). Phosphorylation of
TRIN2 and GATOR2 (Wolfson et al., 2016) and CASTOR1 and IRS1 by S6K inhibits IRS1, thereby forming a negative feedback
GATOR2 (Saxton et al., 2016a); this releases free GATOR2 and loop acting on PI3K and mTORC2 (Shimobayashi and Hall,
thereby activates mTORC1 (Figure 5). SESTRINs may also inhibit 2014). mTORC2 regulates cytoskeletal remodeling, proliferation,
mTORC1 by activating AMPK (Lee et al., 2016). However, and survival by phosphorylating and activating AGC kinase fam-
budding yeast lacks SESTRIN and CASTOR orthologs (Wolfson ily members like AKT at Ser473, PKC (protein kinase C) and SGK
and Sabatini, 2017). Whether, and, if so, how arginine or leucine (serum/glucocorticoid-regulated kinase) (Guri and Hall, 2016).
availability is transmitted to TORC1 in organisms lacking these
proteins is not known. Leucine and glutamine can also activate
mTORC1 via glutaminolysis and consequent production of a-ke- Yin-Yang: Regulation of mTORC1 by AMPK
toglutarate upstream of RAGs (Durán et al., 2013; Durán et al., If the energy status of cells is compromised, it would not be a
2012), while glutamine also activates mTORC1 independently sensible idea for them to grow or divide, even if nutrients were
of the RAGs via the small GTPase ARF1 and the v-ATPase (Jew- still available. It therefore makes sense that AMPK should switch
ell et al., 2015). off mTORC1. Indeed, AMPK activation switches off the mTORC1
It has been reported that LeuRS (leucyl-tRNA synthetase) acts complex by twin mechanisms:
as a cytoplasmic leucine sensor to activate mTORC1 via a RAG- 1. AMPK phosphorylates TSC2 at Thr1271 and Ser1387 (res-
independent mechanism. Leucine-bound LeuRS binds and acti- idue numbering from human isoform 1 (NP_000539); these
vates the class-III phosphoinositide kinase VPS34 that is present sites are referred to as Thr1227 and Ser1345 in the original
in non-autophagic structures. Active VPS34 stimulates PLD1 paper (Inoki et al., 2003). Mutation of these two sites was
(phospholipase D1), thereby increasing phosphatidic acid levels, found to reduce the ability of the glycolytic inhibitor 2-de-
which promote lysosomal activation of mTORC1 (Yoon et al., oxyglucose to inhibit S6K and 4EBP phosphorylation. This
2016; Yoon et al., 2011). phosphorylation is sometimes assumed to promote the
In some cell types, such as epithelial, glial, and mesenchymal GAP activity of the TSC complex toward RHEB, although
stem cells, leucine can activate mTORC1 via production of this has not been directly demonstrated.
acetyl-CoA. Acetyl-CoA stimulates the acetyl transferase 2. AMPK directly phosphorylates the RAPTOR component of
EP300 to acetylate RAPTOR at Lys1097, thereby promoting mTORC1 at two sites, Ser722 and Ser792. Once again,
mTORC1 activity (Son et al., 2019). The acetylated residue is mutation of these two sites was found to reduce the ability
located in the WD40 b-propeller of RAPTOR, close to the ARM of the AMPK activators, AICAR or phenformin, to inhibit
domain (Figure 4). It is unclear whether RAPTOR acetylation af- S6K and 4EBP phosphorylation (Gwinn et al., 2008),
fects mTORC1 structure. although the detailed mechanism for this inhibitory effect
Finally, methionine signals to mTORC1 through synthesis of remains unclear. Ser722 and Ser792 lie in a structurally un-
the methyl donor SAM. SAM availability is transmitted to characterized, and likely disordered, region within the
mTORC1 via SAMTOR (SAM sensor upstream of mTORC1), RAPTOR ARM domain (residues 687–805) (Figure 4)—
with SAM inhibiting the interaction between SAMTOR and note that some publications incorrectly place Ser792 in
GATOR1, thereby activating mTORC1 (Gu et al., 2017a). the RAPTOR b-propeller. Curiously, PKA (cyclic AMP-
Downstream Targets of mTORC1 dependent protein kinase) phosphorylates RAPTOR on
TOR promotes cell growth by stimulating anabolic processes, Ser791, but not Ser792, and is reported to either inhibit
such as ribosome biogenesis and protein, lipid, and nucleotide (Jewell et al., 2019) or activate (Liu et al., 2016)
synthesis, while repressing catabolic processes like autophagy mTORC1—the reasons for this discrepancy are not clear.
(Ben-Sahra and Manning, 2017; Saxton and Sabatini, 2017; Shi-
mobayashi and Hall, 2014). mTORC1 promotes protein synthesis These mechanisms may be at least partly conserved across
by phosphorylating (1) S6K at Thr389 in its hydrophobic motif, to eukaryotes. Although there appear to be no direct orthologs of
increase translation initiation and elongation, and (2) 4EBP, to pro- TSC2 in either budding yeast or plants, there is evidence that
mote cap-dependent translation. mTORC1 also induces purine phosphorylation of the RAPTOR orthologs in S. cerevisiae
synthesis via the tetrahydrofolate cycle (Ben-Sahra et al., 2016) (Hughes Hallett et al., 2015) and plants (Nukarinen et al., 2016)

Cell Metabolism 31, March 3, 2020 483


Cell Metabolism

Review

also leads to inactivation of TORC1 in those organisms. While to requirements are engulfed into lysosomes, where they
these effects were dependent upon the AMPK orthologs, neither are broken down to recycle their components for catabolism or
of the two well-defined sites for AMPK in mammalian RAPTOR re-use. By phosphorylating the autophagy-initiating kinase
(Gwinn et al., 2008) is conserved in S. cerevisiae, and only one ULK1 at distinct sites, AMPK activates while mTORC1 inhibits
is conserved in plants. The detailed mechanisms for these ef- autophagy (Egan et al., 2011b; Kim et al., 2011). AMPK can
fects may therefore be different. therefore promote autophagy not only by direct phosphorylation
These results therefore show that activation of mammalian of ULK1 but also indirectly by inactivating mTORC1 via mecha-
AMPK inhibits mTORC1 via two mechanisms, equivalent to the nisms discussed in the previous paragraph.
fail-safe method of using both ‘‘belt and braces (suspenders)’’ One key downstream target of ULK1 is BECLIN-1, which
to hold up one’s trousers (pants)! A major effect of mTORC1 acti- forms a complex with VPS34, a class-III phosphoinositide kinase
vation is to promote translation, particularly of mRNAs encoding that generates phosphoinositide-3-phosphate (PI3P) on intracel-
proteins required for rapid cell growth, including ribosomal pro- lular membranes. PI3P recruits to those membranes proteins
teins. Since protein synthesis accounts for as much as 20% of containing PI3P-binding domains, which mediate subsequent
total energy turnover in rapidly growing cells (Buttgereit and membrane-trafficking events. VPS34 occurs in several distinct
Brand, 1995), switching it off would have a major effect to complexes; AMPK appears to activate complexes involved in
conserve energy. autophagy by phosphorylating BECLIN-1, while inhibiting those
Although it is therefore clear that AMPK inhibits mTORC1, very involved in other membrane-trafficking events by phosphory-
recently, it has been reported, rather counter-intuitively, that it ac- lating VPS34 itself; this switch depends on the presence of
tivates mTORC2 (Kazyken et al., 2019). Treatment of serum- ATG14L in the former complex (Kim et al., 2013). Thus, AMPK
deprived mouse embryo fibroblasts, HEK293 cells, or primary may divert membrane traffic (an energy-requiring process) to-
mouse hepatocytes with AMPK activators like AICAR, biguanides, ward the autophagy/mitophagy pathway and away from other
or A-769662 was found to increase phosphorylation of the trafficking events that might be a luxury in cells experiencing
mTORC2 site on AKT, Ser473. Although these activators all glucose starvation or energy stress.
have known ‘‘off-target’’ (i.e., AMPK-independent) effects, and As well as their acute effects on autophagy, in the longer term,
more specific AMPK activators are now available, their effects AMPK and mTORC1 also act antagonistically via effects on the
were reduced, although not eliminated, in cells with AMPK related transcription factors EB and E3 (TFEB and TFE3), which
knocked out or knocked down, suggesting that they were at least induce genes involved in lysosome biogenesis and autophagy.
partly mediated by AMPK. The effects were associated with phos- mTORC1 directly phosphorylates TFEB and TFE3, and this pro-
phorylation of Ser1261 on mTOR and unidentified site(s) on RIC- motes their retention in the cytoplasm, inhibiting their transcrip-
TOR, although Ser1261 phosphorylation did not appear to be tional functions (Puertollano et al., 2018). By contrast, AMPK
required for enhanced phosphorylation of AKT. The authors pro- activation promotes dephosphorylation and nuclear transloca-
posed that the activation of mTORC2 by AMPK represents part tion of TFEB, an effect that appears to be at least partially inde-
of the mechanism by which the latter increases cell survival during pendent of mTORC1 (Collodet et al., 2019). One possible mech-
energetic stress and in some circumstances may therefore para- anism for increased transcription at TFEB/TFE3-regulated
doxically promote tumorigenesis (Kazyken et al., 2019). promoters in response to AMPK activation is the increased
In addition, there seems to be a dual ‘‘belt and braces’’ system expression of CARM1 (coactivator-associated arginine methyl-
to turn off mTORC1 when cells are facing shortage of glucose transferase-1) due to downregulation of a E3 ubiquitin ligase
supply. Besides the above-mentioned mechanisms involving containing SKP2 (S-phase kinase-associated protein-2) (Shin
phosphorylation of mTORC1-related factors by AMPK, glucose et al., 2016). Another transcription factor, FOXO3a, is phosphor-
deprivation can inactivate mTORC1 independently of AMPK. Mu- ylated by AMPK at several sites (Greer et al., 2007), and this en-
tations of RAGA/B that abolish GTPase activity completely abro- hances its ability to repress SKP2 expression. The final link in this
gated inhibition of mTORC1 by glucose starvation, despite intact proposed chain of events is that CARM1 is recruited to pro-
activation of AMPK, suggesting that RAGs or RAG-interacting moters of genes involved in autophagy and lysosome biogenesis
partners may play a more direct role in controlling mTORC1 in by TFEB, leading to methylation of Arg17 on histone H3 and
response to nutrients (Efeyan et al., 2013; Kalender et al., 2010). consequent activation of transcription at those sites (Shin
Indeed, in low glucose, AXIN translocates to the surface of the et al., 2016).
lysosome and interacts with the v-ATPase and Ragulator, thereby
facilitating the release of mTORC1 from the lysosomal surface Yang-Yin: Regulation of AMPK by TORC1 and/or
(Zhang et al., 2014). Additional evidence for AMPK-independent Upstream Pathways
regulation is that mTORC1 suppression after glucose starvation There is one report that rapamycin treatment of budding yeast, in
occurs several hours later in AXIN null compared to AXIN wild- wild-type strains but not in strains expressing a TOR1 mutation
type cells in which AMPKa1/a2 had been knocked out (Zhang that confers rapamycin resistance, increases phosphorylation
et al., 2014). This additional device highlights the importance of in- of Thr210 in Snf1 (equivalent to Thr172 in mammalian AMPK)
hibiting mTORC1 when glucose is absent. (Orlova et al., 2006). In mouse embryo fibroblasts, neither rapa-
mycin nor the catalytic site inhibitor of mTOR, Torin1, appeared
Antagonistic Effects of AMPK and mTORC1 on to affect AMPK activity (Zhang et al., 2014). However, while this
Autophagy and Lysosome Biogenesis article was in final preparation, Ling et al. (2020) provided evi-
Autophagy, of which mitophagy (discussed above) is a special dence, using cells from both S. pombe and mammals, that
case, is the process by which cellular contents that are surplus TORC1 directly phosphorylates the AMPK-a1 and a2 subunits

484 Cell Metabolism 31, March 3, 2020


Cell Metabolism

Review

at conserved serine residues within the a-linkers, leading to inhi- phorylation (Scott et al., 2014) and partly because the natural li-
bition of AMPK. The detailed mechanism for this effect remains gands that bind to the ADaM site, if they exist, have not yet been
unclear, because the phosphorylation did not appear to affect identified.
phosphorylation or dephosphorylation of Thr172, or allosteric
activation by AMP, in cell-free assays (Ling et al., 2020). AMPK Conclusions and Perspectives
can also be downregulated by the upstream insulin signaling We have argued in this review that the AMPK and TOR pathways
pathway that activates mTORC1. The insulin-stimulated protein arose very early during eukaryotic evolution and may have been
kinase, AKT, phosphorylates Ser496 (human numbering, required to regulate cell growth in response to the availability of
Q13131) in the a1 catalytic subunit of AMPK (Horman et al., the energy or nutrients provided by some of the newly acquired
2006), and this downregulates (while not completely abolishing) subcellular compartments, such as mitochondria or lysosomes/
AMPK signaling by inhibiting the phosphorylation of Thr172 by vacuoles. The recent findings that lysosomes/vacuoles repre-
LKB1 (Hawley et al., 2014). Ser496 in AMPK-a1 can also be sent key hubs for nutrient sensing by both AMPK and TOR
phosphorylated by PKC (Heathcote et al., 2016) and PKA (Hurley may reflect the fact that early unicellular eukaryotes utilized
et al., 2006). Ser496 occurs in a serine/threonine-rich sequence phagocytosis or pinocytosis for feeding, with nutrients being
just prior to the C-terminal a-helix of AMPK-a1 that has been delivered initially to lysosomes or the vacuole, which, in a unicel-
termed the ‘‘ST loop’’ (Figure 2). A similar sequence is present lular eukaryote, can therefore be considered to be equivalent to
in the a2 isoform, although in that case, the serine residue equiv- the gut. Just as the gut (and associated endocrine pancreas) of
alent to Ser496 (Ser491) is a poor substrate for AKT and appears multicellular animals has become a hub for nutrient sensing
to be modified by autophosphorylation instead (Hawley et al., and signaling, so perhaps did the lysosome or vacuole of unicel-
2014) (it should therefore not be assumed, as is often done, lular eukaryotes.
that the regulation of the two isoforms by ST loop phosphoryla- AMPK can be regarded as representing the Yin or ‘‘dark’’ side
tion is identical). Relevant to this, Ser491 in AMPK-a2 has been of growth control that is activated by lack of energy or nutrients
reported to be phosphorylated by S6K1 (Dagon et al., 2012), and switches off cell growth, while TOR represents the Yang or
which is interesting because the latter is phosphorylated and ‘‘bright’’ side that is activated by availability of nutrients and pro-
activated by mTORC1. However, it is puzzling why there was motes cell growth. In general, TOR pathways promote anabolic
no phosphorylation of Ser491 in the absence of S6K1 in this activities, while AMPK pathways exercise a brake on them.
study (Dagon et al., 2012), when others have observed that These pathways clearly act in opposition to each other, and it
Ser491 in a2 complexes undergoes rapid autophosphorylation is not surprising, as discussed in this review, that there are com-
(Hawley et al., 2014). plex interactions between them. As in Taoist philosophy, the
The ST loop may be subject to multisite phosphorylation, exquisite balance between Yin and Yang ultimately ensures ho-
because GSK3 has been reported to phosphorylate sequentially meostasis and a healthy cell or organism.
within the ST loop of a1 at Thr490, Ser486, and Thr482 (human
numbering, Q13131), which was proposed to promote Thr172 ACKNOWLEDGMENTS
dephosphorylation (Suzuki et al., 2013). Interestingly, the ST
loop is also present in AMPK-a orthologs from C. elegans and M.N.H. acknowledges support from the European Research Council (MERiC,
vertebrates but is absent in those from D. melanogaster and 609883), the Louis Jeantet Foundation, and the Swiss National Science Foun-
dation (310030B-179569). S.C.L. acknowledges supports from the National
S. cerevisiae, suggesting that it is a regulatory sequence that Natural Science Foundation of China (#31730058, #31430094, #31601152
has been inserted during evolution. In the currently available and #31690101) and the National Key R&D Program of China
crystal structures of mammalian heterotrimers, the ST loop has (2016YFA0502001). D.G.H. acknowledges support from an Investigator Award
from the Wellcome Trust (204766). We thank Stefan Imseng (Biozentrum, Uni-
either been deliberately deleted or is not resolved. However, versity of Basel) for assistance in preparation of Figure 4.
the residues at either end of the missing loop lie just 20 and
40 Å from Thr172, suggesting that once phosphorylated, the REFERENCES
loop might interact with the kinase domain and physically block
access to Thr172 (Figure 2). Indeed, there is experimental sup- Alderson, A., Sabelli, P.A., Dickinson, J.R., Cole, D., Richardson, M., Kreis, M.,
port for this model (Hawley et al., 2014). Shewry, P.R., and Halford, N.G. (1991). Complementation of snf1, a mutation
affecting global regulation of carbon metabolism in yeast, by a plant protein ki-
Another potential ‘‘Yang-Yin’’ interaction involves the phos- nase cDNA. Proc. Natl. Acad. Sci. USA 88, 8602–8605.
phorylation of AMPK by ULK1, the autophagy-regulating kinase
Alessi, D.R., Sakamoto, K., and Bayascas, J.R. (2006). LKB1-dependent
that is inactivated/activated by phosphorylation at distinct sites signaling pathways. Annu. Rev. Biochem. 75, 137–163.
by mTORC1/AMPK, respectively (Egan et al., 2011a). ULK1
has been reported to phosphorylate Ser108 on AMPK-b1 but Anandapadamanaban, M., Masson, G.R., Perisic, O., Berndt, A., Kaufman, J.,
Johnson, C.M., Santhanam, B., Rogala, K.B., Sabatini, D.M., and Williams,
not -b2 (Dite et al., 2017). Phosphorylation of Ser-108 is known R.L. (2019). Architecture of human Rag GTPase heterodimers and their com-
to stabilize the ADaM site (see above) by interacting with plex with mTORC1. Science 366, 203–210.
conserved threonine and lysine residues on the N-lobe of the a Anashkin, V.A., Baykov, A.A., and Lahti, R. (2017). Enzymes regulated via cys-
subunit kinase domain (Calabrese et al., 2014; Xiao et al., tathionine beta-synthase domains. Biochemistry (Mosc.) 82, 1079–1087.
2013) and is required for allosteric activation of AMPK by
Apfeld, J., O’Connor, G., McDonagh, T., DiStefano, P.S., and Curtis, R. (2004).
ADaM site ligands both with purified AMPK (Scott et al., 2014) The AMP-activated protein kinase AAK-2 links energy levels and insulin-like
and in intact cells (Dite et al., 2017). However, understanding signals to lifespan in C. elegans. Genes Dev. 18, 3004–3009.
the significance of this requires further study, partly because Aylett, C.H.S., Sauer, E., Imseng, S., Boehringer, D., Hall, M.N., Ban, N., and
Ser108 is also rapidly modified by AMPK itself by cis-autophos- Maier, T. (2016). Architecture of human mTOR complex 1. Science 351, 48–52.

Cell Metabolism 31, March 3, 2020 485


Cell Metabolism

Review
Baena-González, E., and Sheen, J. (2008). Convergent energy and stress Chantranupong, L., Wolfson, R.L., Orozco, J.M., Saxton, R.A., Scaria, S.M.,
signaling. Trends Plant Sci. 13, 474–482. Bar-Peled, L., Spooner, E., Isasa, M., Gygi, S.P., and Sabatini, D.M. (2014).
The Sestrins interact with GATOR2 to negatively regulate the amino-acid-
Baena-González, E., Rolland, F., Thevelein, J.M., and Sheen, J. (2007). A cen- sensing pathway upstream of mTORC1. Cell Rep. 9, 1–8.
tral integrator of transcription networks in plant stress and energy signalling.
Nature 448, 938–942. Chantranupong, L., Scaria, S.M., Saxton, R.A., Gygi, M.P., Shen, K., Wyant,
G.A., Wang, T., Harper, J.W., Gygi, S.P., and Sabatini, D.M. (2016). The
Bar-Peled, L., Schweitzer, L.D., Zoncu, R., and Sabatini, D.M. (2012). Ragula- CASTOR Proteins Are Arginine Sensors for the mTORC1 Pathway. Cell 165,
tor is a GEF for the rag GTPases that signal amino acid levels to mTORC1. Cell 153–164.
150, 1196–1208.
Chao, L.H., and Avruch, J. (2019). Cryo-EM insight into the structure of MTOR
Bar-Peled, L., Chantranupong, L., Cherniack, A.D., Chen, W.W., Ottina, K.A., complex 1 and its interactions with Rheb and substrates. F1000Res. 8, F1000.
Grabiner, B.C., Spear, E.D., Carter, S.L., Meyerson, M., and Sabatini, D.M.
(2013). A Tumor suppressor complex with GAP activity for the Rag GTPases Chen, L., Jiao, Z.H., Zheng, L.S., Zhang, Y.Y., Xie, S.T., Wang, Z.X., and Wu,
that signal amino acid sufficiency to mTORC1. Science 340, 1100–1106. J.W. (2009). Structural insight into the autoinhibition mechanism of AMP-acti-
vated protein kinase. Nature 459, 1146–1149.
, D., Berndt, A., Ohashi, Y., Johnson, C.M., and Williams, R.L. (2016).
Baretic
Tor forms a dimer through an N-terminal helical solenoid with a complex topol- Chen, L., Xin, F.J., Wang, J., Hu, J., Zhang, Y.Y., Wan, S., Cao, L.S., Lu, C., Li,
ogy. Nat. Commun. 7, 11016. P., Yan, S.F., et al. (2013). Conserved regulatory elements in AMPK. Nature
498, E8–E10.
Barnes, K., Ingram, J.C., Porras, O.H., Barros, L.F., Hudson, E.R., Fryer, L.G.,
Chen, X., Liu, M., Tian, Y., Li, J., Qi, Y., Zhao, D., Wu, Z., Huang, M., Wong,
Foufelle, F., Carling, D., Hardie, D.G., and Baldwin, S.A. (2002). Activation of
C.C.L., Wang, H.-W., et al. (2018). Cryo-EM structure of human mTOR com-
GLUT1 by metabolic and osmotic stress: potential involvement of AMP-acti-
plex 2. Cell Research 28, 518–528.
vated protein kinase (AMPK). J. Cell Sci. 115, 2433–2442.
Chia, I.V., and Costantini, F. (2005). Mouse axin and axin2/conductin proteins
Bateman, A. (1997). The structure of a domain common to archaebacteria and are functionally equivalent in vivo. Mol. Cell. Biol. 25, 4371–4376.
the homocystinuria disease protein. Trends Biochem. Sci. 22, 12–13.
Chiu, M.I., Katz, H., and Berlin, V. (1994). RAPT1, a mammalian homolog of
Ben-Sahra, I., and Manning, B.D. (2017). mTORC1 signaling and the metabolic yeast Tor, interacts with the FKBP12/rapamycin complex. Proc. Natl. Acad.
control of cell growth. Curr. Opin. Cell Biol. 45, 72–82. Sci. USA 91, 12574–12578.
Ben-Sahra, I., Howell, J.J., Asara, J.M., and Manning, B.D. (2013). Stimulation Cokorinos, E.C., Delmore, J., Reyes, A.R., Albuquerque, B., Kjobsted, R., Jor-
of de novo pyrimidine synthesis by growth signaling through mTOR and S6K1. gensen, N.O., Tran, J.L., Jatkar, A., Cialdea, K., Esquejo, R.M., et al. (2017).
Science 339, 1323–1328. Activation of skeletal muscle AMPK promotes glucose disposal and glucose
lowering in non-human primates and mice. Cell Metab. 25, 1147–1159.e1110.
Ben-Sahra, I., Hoxhaj, G., Ricoult, S.J.H., Asara, J.M., and Manning, B.D.
(2016). mTORC1 induces purine synthesis through control of the mitochondrial Collodet, C., Foretz, M., Deak, M., Bultot, L., Metairon, S., Viollet, B., Lefebvre,
tetrahydrofolate cycle. Science 351, 728–733. G., Raymond, F., Parisi, A., Civiletto, G., et al. (2019). AMPK promotes induc-
tion of the tumor suppressor FLCN through activation of TFEB independently
Benjamin, D., Colombi, M., Moroni, C., and Hall, M.N. (2011). Rapamycin of mTOR. FASEB J. 33, 12374–12391.
passes the torch: a new generation of mTOR inhibitors. Nat. Rev. Drug Discov.
10, 868–880. Cool, B., Zinker, B., Chiou, W., Kifle, L., Cao, N., Perham, M., Dickinson, R.,
Adler, A., Gagne, G., Iyengar, R., et al. (2006). Identification and characteriza-
Binda, M., Péli-Gulli, M.-P., Bonfils, G., Panchaud, N., Urban, J., Sturgill, T.W., tion of a small molecule AMPK activator that treats key components of type 2
Loewith, R., and De Virgilio, C. (2009). The Vam6 GEF controls TORC1 by acti- diabetes and the metabolic syndrome. Cell Metab. 3, 403–416.
vating the EGO complex. Mol. Cell 35, 563–573.
Dagon, Y., Hur, E., Zheng, B., Wellenstein, K., Cantley, L.C., and Kahn, B.B.
Böhme, U., Otto, T.D., Cotton, J.A., Steinbiss, S., Sanders, M., Oyola, S.O., (2012). p70S6 kinase phosphorylates AMPK on serine 491 to mediate leptin’s
Nicot, A., Gandon, S., Patra, K.P., Herd, C., et al. (2018). Complete avian ma- effect on food intake. Cell Metab. 16, 104–112.
laria parasite genomes reveal features associated with lineage-specific evolu-
tion in birds and mammals. Genome Res. 28, 547–560. Davies, S.P., Helps, N.R., Cohen, P.T., and Hardie, D.G. (1995). 50 -AMP in-
hibits dephosphorylation, as well as promoting phosphorylation, of the AMP-
Brown, E.J., Albers, M.W., Shin, T.B., Ichikawa, K., Keith, C.T., Lane, W.S., and activated protein kinase. Studies using bacterially expressed human protein
Schreiber, S.L. (1994). A mammalian protein targeted by G1-arresting rapamy- phosphatase-2C alpha and native bovine protein phosphatase-2AC. FEBS
cin-receptor complex. Nature 369, 756–758. Lett. 377, 421–425.

Bultot, L., Guigas, B., Von Wilamowitz-Moellendorff, A., Maisin, L., Vertom- de Duve, C. (2005). The lysosome turns fifty. Nat. Cell Biol. 7, 847–849.
men, D., Hussain, N., Beullens, M., Guinovart, J.J., Foretz, M., Viollet, B.,
Demetriades, C., Doumpas, N., and Teleman, A.A. (2014). Regulation of
et al. (2012). AMP-activated protein kinase phosphorylates and inactivates
TORC1 in response to amino acid starvation via lysosomal recruitment of
liver glycogen synthase. Biochem. J. 443, 193–203.
TSC2. Cell 156, 786–799.
Buttgereit, F., and Brand, M.D. (1995). A hierarchy of ATP-consuming pro- Demetriades, C., Plescher, M., and Teleman, A.A. (2016). Lysosomal recruit-
cesses in mammalian cells. Biochem. J. 312, 163–167. ment of TSC2 is a universal response to cellular stress. Nat. Commun.
7, 10662.
Calabrese, M.F., Rajamohan, F., Harris, M.S., Caspers, N.L., Magyar, R.,
Withka, J.M., Wang, H., Borzilleri, K.A., Sahasrabudhe, P.V., Hoth, L.R., DeVit, M.J., and Johnston, M. (1999). The nuclear exportin Msn5 is required for
et al. (2014). Structural basis for AMPK activation: natural and synthetic ligands nuclear export of the Mig1 glucose repressor of Saccharomyces cerevisiae.
regulate kinase activity from opposite poles by different molecular mecha- Curr. Biol. 9, 1231–1241.
nisms. Structure 22, 1161–1172.
Dibble, C.C., Elis, W., Menon, S., Qin, W., Klekota, J., Asara, J.M., Finan, P.M.,
Cantó, C., Gerhart-Hines, Z., Feige, J.N., Lagouge, M., Noriega, L., Milne, J.C., Kwiatkowski, D.J., Murphy, L.O., and Manning, B.D. (2012). TBC1D7 is a third
Elliott, P.J., Puigserver, P., and Auwerx, J. (2009). AMPK regulates energy subunit of the TSC1-TSC2 complex upstream of mTORC1. Mol. Cell 47,
expenditure by modulating NAD+ metabolism and SIRT1 activity. Nature 535–546.
458, 1056–1060.
Dite, T.A., Ling, N.X.Y., Scott, J.W., Hoque, A., Galic, S., Parker, B.L., Ngoei,
Carling, D., Zammit, V.A., and Hardie, D.G. (1987). A common bicyclic protein K.R.W., Langendorf, C.G., O’Brien, M.T., Kundu, M., et al. (2017). The auto-
kinase cascade inactivates the regulatory enzymes of fatty acid and choles- phagy initiator ULK1 sensitizes AMPK to allosteric drugs. Nat. Commun.
terol biosynthesis. FEBS Lett. 223, 217–222. 8, 571.

Celenza, J.L., and Carlson, M. (1986). A yeast gene that is essential for release Ducommun, S., Deak, M., Sumpton, D., Ford, R.J., Núñez Galindo, A., Kuss-
from glucose repression encodes a protein kinase. Science 233, 1175–1180. mann, M., Viollet, B., Steinberg, G.R., Foretz, M., Dayon, L., et al. (2015). Motif

486 Cell Metabolism 31, March 3, 2020


Cell Metabolism

Review
affinity and mass spectrometry proteomic approach for the discovery of directly regulates the mammalian FOXO3 transcription factor. J. Biol. Chem.
cellular AMPK targets: identification of mitochondrial fission factor as a new 282, 30107–30119.
AMPK substrate. Cell. Signal. 27, 978–988.
Gu, X., Orozco, J.M., Saxton, R.A., Condon, K.J., Liu, G.Y., Krawczyk, P.A.,
Dunlop, E.A., and Tee, A.R. (2013). The kinase triad, AMPK, mTORC1 and Scaria, S.M., Harper, J.W., Gygi, S.P., and Sabatini, D.M. (2017a). SAMTOR
ULK1, maintains energy and nutrient homoeostasis. Biochem. Soc. Trans. is an S-adenosylmethionine sensor for the mTORC1 pathway. Science 358,
41, 939–943. 813–818.

Durán, R.V., Oppliger, W., Robitaille, A.M., Heiserich, L., Skendaj, R., Gottlieb, Gu, X., Yan, Y., Novick, S.J., Kovach, A., Goswami, D., Ke, J., Tan, M.H.E.,
E., and Hall, M.N. (2012). Glutaminolysis activates Rag-mTORC1 signaling. Wang, L., Li, X., de Waal, P.W., et al. (2017b). Deconvoluting AMP-activated
Mol. Cell 47, 349–358. protein kinase (AMPK) adenine nucleotide binding and sensing. J. Biol.
Chem. 292, 12653–12666.
Durán, R.V., MacKenzie, E.D., Boulahbel, H., Frezza, C., Heiserich, L., Tardito,
S., Bussolati, O., Rocha, S., Hall, M.N., and Gottlieb, E. (2013). HIF-indepen- Guri, Y., and Hall, M.N. (2016). mTOR Signaling Confers Resistance to Tar-
dent role of prolyl hydroxylases in the cellular response to amino acids. Onco- geted Cancer Drugs. Trends Cancer 2, 688–697.
gene 32, 4549–4556.
Gwinn, D.M., Shackelford, D.B., Egan, D.F., Mihaylova, M.M., Mery, A., Vas-
Efeyan, A., Zoncu, R., Chang, S., Gumper, I., Snitkin, H., Wolfson, R.L., Kirak, quez, D.S., Turk, B.E., and Shaw, R.J. (2008). AMPK phosphorylation of raptor
O., Sabatini, D.D., and Sabatini, D.M. (2013). Regulation of mTORC1 by the mediates a metabolic checkpoint. Mol. Cell 30, 214–226.
Rag GTPases is necessary for neonatal autophagy and survival. Nature 493,
679–683. Hall, M.N. (2013). Talks about TORCs: recent advancesin target of rapamycin
signalling. On mTOR nomenclature. Biochem. Soc. Trans. 41, 887–888.
Egan, D., Kim, J., Shaw, R.J., and Guan, K.L. (2011a). The autophagy initiating
kinase ULK1 is regulated via opposing phosphorylation by AMPK and mTOR. Hara, K., Maruki, Y., Long, X., Yoshino, K., Oshiro, N., Hidayat, S., Tokunaga,
Autophagy 7, 643–644. C., Avruch, J., and Yonezawa, K. (2002). Raptor, a binding partner of target of
rapamycin (TOR), mediates TOR action. Cell 110, 177–189.
Egan, D.F., Shackelford, D.B., Mihaylova, M.M., Gelino, S., Kohnz, R.A., Mair,
W., Vasquez, D.S., Joshi, A., Gwinn, D.M., Taylor, R., et al. (2011b). Phosphor- Hardie, D.G., Carling, D., and Sim, A.T.R. (1989). The AMP-activated protein
ylation of ULK1 (hATG1) by AMP-activated protein kinase connects energy kinase - a multisubstrate regulator of lipid metabolism. Trends Biochem. Sci.
sensing to mitophagy. Science 331, 456–461. 14, 20–23.

Hardie, D.G., Schaffer, B.E., and Brunet, A. (2016). AMPK: an energy-sensing


Ferrer, A., Caelles, C., Massot, N., and Hegardt, F.G. (1985). Activation of rat
pathway with multiple inputs and outputs. Trends Cell Biol. 26, 190–201.
liver cytosolic 3-hydroxy-3-methylglutaryl coenzyme A reductase kinase by
adenosine 50 -monophosphate. Biochem. Biophys. Res. Commun. 132, Hawley, S.A., Selbert, M.A., Goldstein, E.G., Edelman, A.M., Carling, D., and
497–504. Hardie, D.G. (1995). 50 -AMP activates the AMP-activated protein kinase
cascade, and Ca2+/calmodulin activates the calmodulin-dependent protein
Fogarty, S., Ross, F.A., Vara Ciruelos, D., Gray, A., Gowans, G.J., and Hardie,
kinase I cascade, via three independent mechanisms. J. Biol. Chem. 270,
D.G. (2016). AMPK causes cell cycle arrest in LKB1-deficient cells via activa-
27186–27191.
tion of CAMKK2. Mol. Cancer Res. 14, 683–695.
Hawley, S.A., Davison, M., Woods, A., Davies, S.P., Beri, R.K., Carling, D., and
Forte, G.M., Davie, E., Lie, S., Franz-Wachtel, M., Ovens, A.J., Wang, T., Oak-
Hardie, D.G. (1996). Characterization of the AMP-activated protein kinase ki-
ek, B., Hagan, I.M., and Petersen, J. (2019). Import of extracel-
hill, J.S., Mac
nase from rat liver and identification of threonine 172 as the major site at which
lular ATP in yeast and man modulates AMPK and TORC1 signalling. J. Cell
it phosphorylates AMP-activated protein kinase. J. Biol. Chem. 271,
Sci. 132, jcs223925.
27879–27887.
Foster, K.G., Acosta-Jaquez, H.A., Romeo, Y., Ekim, B., Soliman, G.A., Car- €kela
Hawley, S.A., Boudeau, J., Reid, J.L., Mustard, K.J., Udd, L., Ma €, T.P.,
riere, A., Roux, P.P., Ballif, B.A., and Fingar, D.C. (2010). Regulation of Alessi, D.R., and Hardie, D.G. (2003). Complexes between the LKB1 tumor
mTOR complex 1 (mTORC1) by raptor Ser863 and multisite phosphorylation. suppressor, STRAD alpha/beta and MO25 alpha/beta are upstream kinases
J. Biol. Chem. 285, 80–94. in the AMP-activated protein kinase cascade. J. Biol. 2, 28.
Fu, X., Wan, S., Lyu, Y.L., Liu, L.F., and Qi, H. (2008). Etoposide induces ATM- Hawley, S.A., Pan, D.A., Mustard, K.J., Ross, L., Bain, J., Edelman, A.M., Fren-
dependent mitochondrial biogenesis through AMPK activation. PLoS One guelli, B.G., and Hardie, D.G. (2005). Calmodulin-dependent protein kinase ki-
3, e2009. nase-beta is an alternative upstream kinase for AMP-activated protein kinase.
Cell Metab. 2, 9–19.
Fullerton, M.D., Galic, S., Marcinko, K., Sikkema, S., Pulinilkunnil, T., Chen,
Z.P., O’Neill, H.M., Ford, R.J., Palanivel, R., O’Brien, M., et al. (2013). Single Hawley, S.A., Fullerton, M.D., Ross, F.A., Schertzer, J.D., Chevtzoff, C.,
phosphorylation sites in Acc1 and Acc2 regulate lipid homeostasis and the in- Walker, K.J., Peggie, M.W., Zibrova, D., Green, K.A., Mustard, K.J., et al.
sulin-sensitizing effects of metformin. Nat. Med. 19, 1649–1654. (2012). The ancient drug salicylate directly activates AMP-activated protein ki-
nase. Science 336, 918–922.
Gaubitz, C., Oliveira, T.M., Prouteau, M., Leitner, A., Karuppasamy, M., Kon-
stantinidou, G., Rispal, D., Eltschinger, S., Robinson, G.C., Thore, S., et al. Hawley, S.A., Ross, F.A., Gowans, G.J., Tibarewal, P., Leslie, N.R., and Hardie,
(2015). Molecular Basis of the Rapamycin Insensitivity of Target Of Rapamycin D.G. (2014). Phosphorylation by Akt within the ST loop of AMPK-a1 down-reg-
Complex 2. Mol. Cell 58, 977–988. ulates its activation in tumour cells. Biochem. J. 459, 275–287.
González, A., and Hall, M.N. (2017). Nutrient sensing and TOR signaling in Heathcote, H.R., Mancini, S.J., Strembitska, A., Jamal, K., Reihill, J.A., Palmer,
yeast and mammals. EMBO J. 36, 397–408. T.M., Gould, G.W., and Salt, I.P. (2016). Protein kinase C phosphorylates AMP-
activated protein kinase a1 Ser487. Biochem. J. 473, 4681–4697.
Göransson, O., McBride, A., Hawley, S.A., Ross, F.A., Shpiro, N., Foretz, M.,
Viollet, B., Hardie, D.G., and Sakamoto, K. (2007). Mechanism of action of Hedbacker, K., and Carlson, M. (2008). SNF1/AMPK pathways in yeast. Front.
A-769662, a valuable tool for activation of AMP-activated protein kinase. Biosci. 13, 2408–2420.
J. Biol. Chem. 282, 32549–32560.
Heitman, J., Movva, N.R., and Hall, M.N. (1991). Targets for cell cycle arrest by
Gouw, M., Michael, S., Sámano-Sánchez, H., Kumar, M., Zeke, A., Lang, B., the immunosuppressant rapamycin in yeast. Science 253, 905–909.
Bely, B., Chemes, L.B., Davey, N.E., Deng, Z., et al. (2018). The eukaryotic
linear motif resource - 2018 update. Nucleic Acids Res. 46 (D1), D428–D434. Herrero-Martı́n, G., Høyer-Hansen, M., Garcı́a-Garcı́a, C., Fumarola, C., Far-
€a
kas, T., López-Rivas, A., and Ja €ttela
€, M. (2009). TAK1 activates AMPK-
Gowans, G.J., Hawley, S.A., Ross, F.A., and Hardie, D.G. (2013). AMP is a true dependent cytoprotective autophagy in TRAIL-treated epithelial cells. EMBO
physiological regulator of AMP-activated protein kinase by both allosteric acti- J. 28, 677–685.
vation and enhancing net phosphorylation. Cell Metab. 18, 556–566.
Hoppe, S., Bierhoff, H., Cado, I., Weber, A., Tiebe, M., Grummt, I., and Voit, R.
Greer, E.L., Oskoui, P.R., Banko, M.R., Maniar, J.M., Gygi, M.P., Gygi, S.P., (2009). AMP-activated protein kinase adapts rRNA synthesis to cellular energy
and Brunet, A. (2007). The energy sensor AMP-activated protein kinase supply. Proc. Natl. Acad. Sci. USA 106, 17781–17786.

Cell Metabolism 31, March 3, 2020 487


Cell Metabolism

Review
Horman, S., Vertommen, D., Heath, R., Neumann, D., Mouton, V., Woods, A., Karuppasamy, M., Kusmider, B., Oliveira, T.M., Gaubitz, C., Prouteau, M.,
Schlattner, U., Wallimann, T., Carling, D., Hue, L., and Rider, M.H. (2006). Insu- Loewith, R., and Schaffitzel, C. (2017). Cryo-EM structure of Saccharomyces
lin antagonizes ischemia-induced Thr172 phosphorylation of AMP-activated cerevisiae target of rapamycin complex 2. Nat. Commun. 8, 1729.
protein kinase alpha-subunits in heart via hierarchical phosphorylation of
Ser485/491. J. Biol. Chem. 281, 5335–5340. Kazyken, D., Magnuson, B., Bodur, C., Acosta-Jaquez, H.A., Zhang, D., Tong,
X., Barnes, T.M., Steinl, G.K., Patterson, N.E., Altheim, C.H., et al. (2019).
Hudson, E.R., Pan, D.A., James, J., Lucocq, J.M., Hawley, S.A., Green, K.A., AMPK directly activates mTORC2 to promote cell survival during acute ener-
Baba, O., Terashima, T., and Hardie, D.G. (2003). A novel domain in AMP-acti- getic stress. Sci. Signal. 12, eaav3249.
vated protein kinase causes glycogen storage bodies similar to those seen in
hereditary cardiac arrhythmias. Curr. Biol. 13, 861–866. Keith, C.T., and Schreiber, S.L. (1995). PIK-related kinases: DNA repair,
recombination, and cell cycle checkpoints. Science 270, 50–51.
Hughes Hallett, J.E., Luo, X., and Capaldi, A.P. (2015). Snf1/AMPK promotes
the formation of Kog1/Raptor-bodies to increase the activation threshold of Kim, J., and Guan, K.-L. (2019). mTOR as a central hub of nutrient signalling
TORC1 in budding yeast. Elife 4, e09181. and cell growth. Nat. Cell Biol. 21, 63–71.

Hurley, R.L., Anderson, K.A., Franzone, J.M., Kemp, B.E., Means, A.R., and Kim, D.-H., Sarbassov, D.D., Ali, S.M., King, J.E., Latek, R.R., Erdjument-
Witters, L.A. (2005). The Ca2+/calmodulin-dependent protein kinase kinases Bromage, H., Tempst, P., and Sabatini, D.M. (2002). mTOR interacts with
are AMP-activated protein kinase kinases. J. Biol. Chem. 280, 29060–29066. raptor to form a nutrient-sensitive complex that signals to the cell growth ma-
chinery. Cell 110, 163–175.
Hurley, R.L., Barré, L.K., Wood, S.D., Anderson, K.A., Kemp, B.E., Means,
A.R., and Witters, L.A. (2006). Regulation of AMP-activated protein kinase Kim, J., Kundu, M., Viollet, B., and Guan, K.L. (2011). AMPK and mTOR regu-
by multisite phosphorylation in response to agents that elevate cellular late autophagy through direct phosphorylation of Ulk1. Nat. Cell Biol. 13,
cAMP. J. Biol. Chem. 281, 36662–36672. 132–141.

Imamura, K., Ogura, T., Kishimoto, A., Kaminishi, M., and Esumi, H. (2001). Cell Kim, J., Kim, Y.C., Fang, C., Russell, R.C., Kim, J.H., Fan, W., Liu, R., Zhong,
cycle regulation via p53 phosphorylation by a 50 -AMP activated protein kinase Q., and Guan, K.L. (2013). Differential regulation of distinct Vps34 complexes
activator, 5-aminoimidazole- 4-carboxamide-1-beta-D-ribofuranoside, in a by AMPK in nutrient stress and autophagy. Cell 152, 290–303.
human hepatocellular carcinoma cell line. Biochem. Biophys. Res. Commun.
287, 562–567. Kim, J.S., Ro, S.-H., Kim, M., Park, H.-W., Semple, I.A., Park, H., Cho, U.-S.,
Wang, W., Guan, K.-L., Karin, M., and Lee, J.H. (2015). Sestrin2 inhibits
Imseng, S., Aylett, C.H.S., and Maier, T. (2018). Architecture and activation of mTORC1 through modulation of GATOR complexes. Sci. Rep. 5, 9502.
phosphatidylinositol 3-kinase related kinases. Curr. Opin. Struct. Biol. 49,
177–189. Kogan, K., Spear, E.D., Kaiser, C.A., and Fass, D. (2010). Structural conserva-
tion of components in the amino acid sensing branch of the TOR pathway in
Inoki, K., Li, Y., Zhu, T., Wu, J., and Guan, K.-L. (2002). TSC2 is phosphorylated yeast and mammals. J. Mol. Biol. 402, 388–398.
and inhibited by Akt and suppresses mTOR signalling. Nat. Cell Biol. 4,
648–657. Kunz, J., Henriquez, R., Schneider, U., Deuter-Reinhard, M., Movva, N.R., and
Hall, M.N. (1993). Target of rapamycin in yeast, TOR2, is an essential phospha-
Inoki, K., Zhu, T., and Guan, K.L. (2003). TSC2 mediates cellular energy tidylinositol kinase homolog required for G1 progression. Cell 73, 585–596.
response to control cell growth and survival. Cell 115, 577–590.
Kurth-Kraczek, E.J., Hirshman, M.F., Goodyear, L.J., and Winder, W.W.
Jadeja, R.N., Chu, X., Wood, C., Bartoli, M., and Khurana, S. (2019). M3 (1999). 50 AMP-activated protein kinase activation causes GLUT4 transloca-
Muscarinic receptor activation reduces hepatocyte lipid accumulation via tion in skeletal muscle. Diabetes 48, 1667–1671.
CaMKKbeta/AMPK pathway. Biochem. Pharmacol. 169, 113613.
Lane, N. (2006). Power, sex and suicide: mitochondria and the meaning of life
€ger, S., Handschin, C., St-Pierre, J., and Spiegelman, B.M. (2007). AMP-
Ja (Oxford University Press).
activated protein kinase (AMPK) action in skeletal muscle via direct phosphor-
ylation of PGC-1alpha. Proc. Natl. Acad. Sci. USA 104, 12017–12022. Lane, N., and Martin, W. (2010). The energetics of genome complexity. Nature
467, 929–934.
Jaleel, M., McBride, A., Lizcano, J.M., Deak, M., Toth, R., Morrice, N.A., and
Alessi, D.R. (2005). Identification of the sucrose non-fermenting related kinase Langendorf, C.G., and Kemp, B.E. (2015). Choreography of AMPK activation.
SNRK, as a novel LKB1 substrate. FEBS Lett. 579, 1417–1423. Cell Res. 25, 5–6.

Jewell, J.L., Kim, Y.C., Russell, R.C., Yu, F.X., Park, H.W., Plouffe, S.W., Ta- Lantier, L., Fentz, J., Mounier, R., Leclerc, J., Treebak, J.T., Pehmøller, C.,
gliabracci, V.S., and Guan, K.L. (2015). Metabolism. Differential regulation of Sanz, N., Sakakibara, I., Saint-Amand, E., Rimbaud, S., et al. (2014). AMPK
mTORC1 by leucine and glutamine. Science 347, 194–198. controls exercise endurance, mitochondrial oxidative capacity, and skeletal
muscle integrity. FASEB J. 28, 3211–3224.
Jewell, J.L., Fu, V., Hong, A.W., Yu, F.-X., Meng, D., Melick, C.H., Wang, H.,
Lam, W.M., Yuan, H.-X., Taylor, S.S., and Guan, K.L. (2019). GPCR signaling Lee, J.H., Cho, U.-S., and Karin, M. (2016). Sestrin regulation of TORC1: Is
inhibits mTORC1 via PKA phosphorylation of Raptor. Elife 8, e43038. Sestrin a leucine sensor? Sci. Signal. 9, re5.

Ji, H., Ramsey, M.R., Hayes, D.N., Fan, C., McNamara, K., Kozlowski, P., Tor- Levine, T.P., Daniels, R.D., Wong, L.H., Gatta, A.T., Gerondopoulos, A., and
rice, C., Wu, M.C., Shimamura, T., Perera, S.A., et al. (2007). LKB1 modulates Barr, F.A. (2013). Discovery of new Longin and Roadblock domains that
lung cancer differentiation and metastasis. Nature 448, 807–810. form platforms for small GTPases in Ragulator and TRAPP-II. Small GTPases
4, 62–69.
Johanns, M., Pyr Dit Ruys, S., Houddane, A., Vertommen, D., Herinckx, G.,
Hue, L., Proud, C.G., and Rider, M.H. (2017). Direct and indirect activation of Li, X., Wang, L., Zhou, X.E., Ke, J., de Waal, P.W., Gu, X., Tan, M.H., Wang, D.,
eukaryotic elongation factor 2 kinase by AMP-activated protein kinase. Cell. Wu, D., Xu, H.E., and Melcher, K. (2015). Structural basis of AMPK regulation
Signal. 36, 212–221. by adenine nucleotides and glycogen. Cell Res. 25, 50–66.

Jørgensen, S.B., Nielsen, J.N., Birk, J.B., Olsen, G.S., Viollet, B., Andreelli, F., Li, M., Zhang, C.S., Zong, Y., Feng, J.W., Ma, T., Hu, M., Lin, Z., Li, X., Xie, C.,
Schjerling, P., Vaulont, S., Hardie, D.G., Hansen, B.F., et al. (2004). The alpha2- Wu, Y., et al. (2019a). Transient Receptor Potential V Channels Are Essential
5’AMP-activated protein kinase is a site 2 glycogen synthase kinase in skeletal for Glucose Sensing by Aldolase and AMPK. Cell Metab. 30, 508–524.e12.
muscle and is responsive to glucose loading. Diabetes 53, 3074–3081.
Li, S., Lavagnino, Z., Lemacon, D., Kong, L., Ustione, A., Ng, X., Zhang, Y.,
Jung, J., Genau, H.M., and Behrends, C. (2015). Amino Acid-Dependent Wang, Y., Zheng, B., Piwnica-Worms, H., et al. (2019b). Ca(2+)-stimulated
mTORC1 Regulation by the Lysosomal Membrane Protein SLC38A9. Mol. AMPK-dependent phosphorylation of Exo1 protects stressed replication forks
Cell. Biol. 35, 2479–2494. from aberrant resection. Mol. Cell 74, 1123–1137.e1126.

Kalender, A., Selvaraj, A., Kim, S.Y., Gulati, P., Brûlé, S., Viollet, B., Kemp, B.E., Ling, N.X.Y., Kaczmarek, A., Hoque, A., Davie, E., Ngoei, K.R.W., Morrison,
Bardeesy, N., Dennis, P., Schlager, J.J., et al. (2010). Metformin, independent K.R., Smiles, W.J., Forte, G.M., Wang, T., Lie, S., et al. (2020). Nat. Metab.
of AMPK, inhibits mTORC1 in a rag GTPase-dependent manner. Cell Metab. mTORC1 directly inhibits AMPK to promote cell proliferation under nutrient
11, 390–401. stress 2, 41–49.

488 Cell Metabolism 31, March 3, 2020


Cell Metabolism

Review
Liu, D., Bordicchia, M., Zhang, C., Fang, H., Wei, W., Li, J.-L., Guilherme, A., (2012). Mutations in the Arabidopsis homolog of LST8/GbetaL, a partner of the
Guntur, K., Czech, M.P., and Collins, S. (2016). Activation of mTORC1 is target of Rapamycin kinase, impair plant growth, flowering, and metabolic
essential for b-adrenergic stimulation of adipose browning. J. Clin. Invest. adaptation to long days. Plant Cell 24, 463–481.
126, 1704–1716.
Morris, A.J., and Tolan, D.R. (1993). Site-directed mutagenesis identifies
Lizcano, J.M., Göransson, O., Toth, R., Deak, M., Morrice, N.A., Boudeau, J., aspartate 33 as a previously unidentified critical residue in the catalytic mech-
Hawley, S.A., Udd, L., Ma €kela
€, T.P., Hardie, D.G., and Alessi, D.R. (2004). anism of rabbit aldolase A. J. Biol. Chem. 268, 1095–1100.
LKB1 is a master kinase that activates 13 kinases of the AMPK subfamily,
including MARK/PAR-1. EMBO J. 23, 833–843. Myers, R.W., Guan, H.P., Ehrhart, J., Petrov, A., Prahalada, S., Tozzo, E.,
Yang, X., Kurtz, M.M., Trujillo, M., Gonzalez Trotter, D., et al. (2017). Systemic
Loewith, R., Jacinto, E., Wullschleger, S., Lorberg, A., Crespo, J.L., Bonenfant, pan-AMPK activator MK-8722 improves glucose homeostasis but induces
D., Oppliger, W., Jenoe, P., and Hall, M.N. (2002). Two TOR complexes, only cardiac hypertrophy. Science 357, 507–511.
one of which is rapamycin sensitive, have distinct roles in cell growth control.
Mol. Cell 10, 457–468. Nada, S., Hondo, A., Kasai, A., Koike, M., Saito, K., Uchiyama, Y., and Okada,
M. (2009). The novel lipid raft adaptor p18 controls endosome dynamics by
Loh, K., Tam, S., Murray-Segal, L., Huynh, K., Meikle, P.J., Scott, J.W., van anchoring the MEK-ERK pathway to late endosomes. EMBO J. 28, 477–489.
Denderen, B., Chen, Z., Steel, R., LeBlond, N.D., et al. (2018). Inhibition of
Adenosine Monophosphate-activated Protein Kinase-3-Hydroxy-3-Methyl- Ngoei, K.R.W., Langendorf, C.G., Ling, N.X.Y., Hoque, A., Varghese, S., Ca-
glutaryl Coenzyme A Reductase signaling leads to hypercholesterolemia and merino, M.A., Walker, S.R., Bozikis, Y.E., Dite, T.A., Ovens, A.J., et al.
promotes hepatic steatosis and insulin resistance. Hepatol. Commun. (2018). Structural determinants for small-molecule activation of skeletal mus-
3, 84–98. cle AMPK alpha2beta2gamma1 by the glucose importagog SC4. Cell Chem.
Biol. 25, 728–737.e9.
Long, X., Lin, Y., Ortiz-Vega, S., Yonezawa, K., and Avruch, J. (2005). Rheb
binds and regulates the mTOR kinase. Curr. Biol. 15, 702–713. Nicastro, R., Sardu, A., Panchaud, N., and De Virgilio, C. (2017). The Architec-
ture of the Rag GTPase Signaling Network. Biomolecules 7, 48.
Lumbreras, V., Alba, M.M., Kleinow, T., Koncz, C., and Pagès, M. (2001).
Domain fusion between SNF1-related kinase subunits during plant evolution. Nojima, H., Tokunaga, C., Eguchi, S., Oshiro, N., Hidayat, S., Yoshino, K.,
EMBO Rep. 2, 55–60. Hara, K., Tanaka, N., Avruch, J., and Yonezawa, K. (2003). The mammalian
target of rapamycin (mTOR) partner, raptor, binds the mTOR substrates p70
Mackintosh, R.W., Davies, S.P., Clarke, P.R., Weekes, J., Gillespie, J.G., Gibb, S6 kinase and 4E-BP1 through their TOR signaling (TOS) motif. J. Biol.
B.J., and Hardie, D.G. (1992). Evidence for a protein kinase cascade in higher Chem. 278, 15461–15464.
plants. 3-Hydroxy-3-methylglutaryl-CoA reductase kinase. Eur. J. Biochem.
209, 923–931. Nukarinen, E., Na€gele, T., Pedrotti, L., Wurzinger, B., Mair, A., Landgraf, R.,
Börnke, F., Hanson, J., Teige, M., Baena-Gonzalez, E., et al. (2016). Quantita-
Marsin, A.S., Bertrand, L., Rider, M.H., Deprez, J., Beauloye, C., Vincent, M.F., tive phosphoproteomics reveals the role of the AMPK plant ortholog SnRK1 as
Van den Berghe, G., Carling, D., and Hue, L. (2000). Phosphorylation and acti- a metabolic master regulator under energy deprivation. Sci. Rep. 6, 31697.
vation of heart PFK-2 by AMPK has a role in the stimulation of glycolysis during
ischaemia. Curr. Biol. 10, 1247–1255. O’Donnell, A.F., and Schmidt, M.C. (2019). AMPK-mediated regulation of
alpha-arrestins and protein trafficking. Int. J. Mol. Sci. 20, E515.
Marsin, A.S., Bouzin, C., Bertrand, L., and Hue, L. (2002). The stimulation of
glycolysis by hypoxia in activated monocytes is mediated by AMP-activated O’Neill, H.M., Maarbjerg, S.J., Crane, J.D., Jeppesen, J., Jørgensen, S.B.,
protein kinase and inducible 6-phosphofructo-2-kinase. J. Biol. Chem. 277, Schertzer, J.D., Shyroka, O., Kiens, B., van Denderen, B.J., Tarnopolsky,
30778–30783. M.A., et al. (2011). AMP-activated protein kinase (AMPK) beta1beta2 muscle
null mice reveal an essential role for AMPK in maintaining mitochondrial con-
Martina, J.A., Chen, Y., Gucek, M., and Puertollano, R. (2012). MTORC1 func- tent and glucose uptake during exercise. Proc. Natl. Acad. Sci. USA 108,
tions as a transcriptional regulator of autophagy by preventing nuclear trans- 16092–16097.
port of TFEB. Autophagy 8, 903–914.
Oakhill, J.S., Steel, R., Chen, Z.P., Scott, J.W., Ling, N., Tam, S., and Kemp,
Mayer, F.V., Heath, R., Underwood, E., Sanders, M.J., Carmena, D., McCart- B.E. (2011). AMPK is a direct adenylate charge-regulated protein kinase. Sci-
ney, R.R., Leiper, F.C., Xiao, B., Jing, C., Walker, P.A., et al. (2011). ADP reg- ence 332, 1433–1435.
ulates SNF1, the Saccharomyces cerevisiae homolog of AMP-activated pro-
tein kinase. Cell Metab. 14, 707–714. Orlova, M., Kanter, E., Krakovich, D., and Kuchin, S. (2006). Nitrogen availabil-
ity and TOR regulate the Snf1 protein kinase in Saccharomyces cerevisiae. Eu-
McBride, A., Ghilagaber, S., Nikolaev, A., and Hardie, D.G. (2009). The karyot. Cell 5, 1831–1837.
glycogen-binding domain on the AMPK beta subunit allows the kinase to act
as a glycogen sensor. Cell Metab. 9, 23–34. Palm, W., and Thompson, C.B. (2017). Nutrient acquisition strategies of
mammalian cells. Nature 546, 234–242.
Menon, S., Dibble, C.C., Talbott, G., Hoxhaj, G., Valvezan, A.J., Takahashi, H.,
Cantley, L.C., and Manning, B.D. (2014). Spatial control of the TSC complex Pan, D.A., and Hardie, D.G. (2002). A homologue of AMP-activated protein ki-
integrates insulin and nutrient regulation of mTORC1 at the lysosome. Cell nase in Drosophila melanogaster is sensitive to AMP and is activated by ATP
156, 771–785. depletion. Biochem. J. 367, 179–186.

Merlin, J., Evans, B.A., Csikasz, R.I., Bengtsson, T., Summers, R.J., and Panchaud, N., Péli-Gulli, M.-P., and De Virgilio, C. (2013a). Amino acid depri-
Hutchinson, D.S. (2010). The M3-muscarinic acetylcholine receptor stimulates vation inhibits TORC1 through a GTPase-activating protein complex for the
glucose uptake in L6 skeletal muscle cells by a CaMKK-AMPK-dependent Rag family GTPase Gtr1. Sci. Signal. 6, ra42.
mechanism. Cell. Signal. 22, 1104–1113.
Panchaud, N., Péli-Gulli, M.-P., and De Virgilio, C. (2013b). SEACing the GAP
Miranda-Saavedra, D., Stark, M.J., Packer, J.C., Vivares, C.P., Doerig, C., and that nEGOCiates TORC1 activation: evolutionary conservation of Rag GTPase
Barton, G.J. (2007). The complement of protein kinases of the microsporidium regulation. Cell Cycle 12, 2948–2952.
Encephalitozoon cuniculi in relation to those of Saccharomyces cerevisiae and
Schizosaccharomyces pombe. BMC Genomics 8, 309. Papamichos-Chronakis, M., Gligoris, T., and Tzamarias, D. (2004). The Snf1 ki-
nase controls glucose repression in yeast by modulating interactions between
Mitchelhill, K.I., Stapleton, D., Gao, G., House, C., Michell, B., Katsis, F., Wit- the Mig1 repressor and the Cyc8-Tup1 co-repressor. EMBO Rep. 5, 368–372.
ters, L.A., and Kemp, B.E. (1994). Mammalian AMP-activated protein kinase
shares structural and functional homology with the catalytic domain of yeast Parmigiani, A., Nourbakhsh, A., Ding, B., Wang, W., Kim, Y.C., Akopiants, K.,
Snf1 protein kinase. J. Biol. Chem. 269, 2361–2364. Guan, K.-L., Karin, M., and Budanov, A.V. (2014). Sestrins inhibit mTORC1 ki-
nase activation through the GATOR complex. Cell Rep. 9, 1281–1291.
Momcilovic, M., Hong, S.P., and Carlson, M. (2006). Mammalian TAK1 acti-
vates Snf1 protein kinase in yeast and phosphorylates AMP-activated protein Pehmøller, C., Treebak, J.T., Birk, J.B., Chen, S., Mackintosh, C., Hardie, D.G.,
kinase in vitro. J. Biol. Chem. 281, 25336–25343. Richter, E.A., and Wojtaszewski, J.F. (2009). Genetic disruption of AMPK
signaling abolishes both contraction- and insulin-stimulated TBC1D1 phos-
Moreau, M., Azzopardi, M., Clement, G., Dobrenel, T., Marchive, C., Renne, C., phorylation and 14-3-3 binding in mouse skeletal muscle. Am. J. Physiol. En-
Martin-Magniette, M.L., Taconnat, L., Renou, J.P., Robaglia, C., and Meyer, C. docrinol. Metab. 297, E665–E675.

Cell Metabolism 31, March 3, 2020 489


Cell Metabolism

Review
Pelosse, M., Cottet-Rousselle, C., Bidan, C.M., Dupont, A., Gupta, K., Berger, Sancak, Y., Bar-Peled, L., Zoncu, R., Markhard, A.L., Nada, S., and Sabatini,
I., and Schlattner, U. (2019). Synthetic energy sensor AMPfret deciphers ad- D.M. (2010). Ragulator-Rag complex targets mTORC1 to the lysosomal sur-
enylate-dependent AMPK activation mechanism. Nat. Commun. 10, 1038. face and is necessary for its activation by amino acids. Cell 141, 290–303.

Peng, M., Yin, N., and Li, M.O. (2017). SZT2 dictates GATOR control of Sanchez-Cespedes, M., Parrella, P., Esteller, M., Nomoto, S., Trink, B., Eng-
mTORC1 signalling. Nature 543, 433–437. les, J.M., Westra, W.H., Herman, J.G., and Sidransky, D. (2002). Inactivation
of LKB1/STK11 is a common event in adenocarcinomas of the lung. Cancer
Petit, C.S., Roczniak-Ferguson, A., and Ferguson, S.M. (2013). Recruitment of Res. 62, 3659–3662.
folliculin to lysosomes supports the amino acid-dependent activation of Rag
GTPases. J. Cell Biol. 202, 1107–1122. Sanders, M.J., Ali, Z.S., Hegarty, B.D., Heath, R., Snowden, M.A., and Carling,
D. (2007). Defining the mechanism of activation of AMP-activated protein ki-
Polekhina, G., Gupta, A., Michell, B.J., van Denderen, B., Murthy, S., Feil, S.C., nase by the small molecule A-769662, a member of the thienopyridone family.
Jennings, I.G., Campbell, D.J., Witters, L.A., Parker, M.W., et al. (2003). AMPK
J. Biol. Chem. 282, 32539–32548.
beta subunit targets metabolic stress sensing to glycogen. Curr. Biol. 13,
867–871.
Sanli, T., Rashid, A., Liu, C., Harding, S., Bristow, R.G., Cutz, J.C., Singh, G.,
Polekhina, G., Gupta, A., van Denderen, B.J., Feil, S.C., Kemp, B.E., Stapleton, Wright, J., and Tsakiridis, T. (2010). Ionizing radiation activates AMP-activated
D., and Parker, M.W. (2005). Structural basis for glycogen recognition by AMP- kinase (AMPK): a target for radiosensitization of human cancer cells. Int. J. Ra-
activated protein kinase. Structure 13, 1453–1462. diat. Oncol. Biol. Phys. 78, 221–229.

Powis, K., Zhang, T., Panchaud, N., Wang, R., De Virgilio, C., and Ding, J. Saxton, R.A., and Sabatini, D.M. (2017). mTOR Signaling in Growth, Meta-
(2015). Crystal structure of the Ego1-Ego2-Ego3 complex and its role in pro- bolism, and Disease. Cell 168, 960–976.
moting Rag GTPase-dependent TORC1 signaling. Cell Res. 25, 1043–1059.
Saxton, R.A., Chantranupong, L., Knockenhauer, K.E., Schwartz, T.U., and
Prouteau, M., Desfosses, A., Sieben, C., Bourgoint, C., Lydia Mozaffari, N., Sabatini, D.M. (2016a). Mechanism of arginine sensing by CASTOR1 upstream
Demurtas, D., Mitra, A.K., Guichard, P., Manley, S., and Loewith, R. (2017). of mTORC1. Nature 536, 229–233.
TORC1 organized in inhibited domains (TOROIDs) regulate TORC1 activity.
Nature 550, 265–269. Saxton, R.A., Knockenhauer, K.E., Schwartz, T.U., and Sabatini, D.M. (2016b).
The apo-structure of the leucine sensor Sestrin2 is still elusive. Sci. Signal.
Puertollano, R., Ferguson, S.M., Brugarolas, J., and Ballabio, A. (2018). The 9, ra92.
complex relationship between TFEB transcription factor phosphorylation
and subcellular localization. EMBO J. 37, e98804. Saxton, R.A., Knockenhauer, K.E., Wolfson, R.L., Chantranupong, L., Pacold,
M.E., Wang, T., Schwartz, T.U., and Sabatini, D.M. (2016c). Structural basis for
Qian, X., Li, X., Tan, L., Lee, J.H., Xia, Y., Cai, Q., Zheng, Y., Wang, H., Lorenzi, leucine sensing by the Sestrin2-mTORC1 pathway. Science 351, 53–58.
P.L., and Lu, Z. (2018). Conversion of PRPS hexamer to monomer by AMPK-
mediated phosphorylation inhibits nucleotide synthesis in response to energy Schaffer, B.E., Levin, R.S., Hertz, N.T., Maures, T.J., Schoof, M.L., Hollstein,
stress. Cancer Discov. 8, 94–107. P.E., Benayoun, B.A., Banko, M.R., Shaw, R.J., Shokat, K.M., and Brunet, A.
(2015). Identification of AMPK phosphorylation sites reveals a network of pro-
Rebsamen, M., Pochini, L., Stasyk, T., de Araújo, M.E.G., Galluccio, M., Kan- teins involved in cell invasion and facilitates large-scale substrate prediction.
dasamy, R.K., Snijder, B., Fauster, A., Rudashevskaya, E.L., Bruckner, M., Cell Metab. 22, 907–921.
et al. (2015). SLC38A9 is a component of the lysosomal amino acid sensing
machinery that controls mTORC1. Nature 519, 477–481.
Schalm, S.S., and Blenis, J. (2002). Identification of a conserved motif required
Reihill, J.A., Ewart, M.A., Hardie, D.G., and Salt, I.P. (2007). AMP-activated for mTOR signaling. Curr. Biol. 12, 632–639.
protein kinase mediates VEGF-stimulated endothelial NO production. Bio-
chem. Biophys. Res. Commun. 354, 1084–1088. Schalm, S.S., Fingar, D.C., Sabatini, D.M., and Blenis, J. (2003). TOS motif-
mediated raptor binding regulates 4E-BP1 multisite phosphorylation and func-
Robitaille, A.M., Christen, S., Shimobayashi, M., Cornu, M., Fava, L.L., Moes, tion. Curr. Biol. 13, 797–806.
S., Prescianotto-Baschong, C., Sauer, U., Jenoe, P., and Hall, M.N. (2013).
Quantitative phosphoproteomics reveal mTORC1 activates de novo pyrimi- Scott, J.W., Hawley, S.A., Green, K.A., Anis, M., Stewart, G., Scullion, G.A.,
dine synthesis. Science 339, 1320–1323. Norman, D.G., and Hardie, D.G. (2004). CBS domains form energy-sensing
modules whose binding of adenosine ligands is disrupted by disease muta-
Roczniak-Ferguson, A., Petit, C.S., Froehlich, F., Qian, S., Ky, J., Angarola, B., tions. J. Clin. Invest. 113, 274–284.
Walther, T.C., and Ferguson, S.M. (2012). The transcription factor TFEB links
mTORC1 signaling to transcriptional control of lysosome homeostasis. Sci. Scott, J.W., Ling, N., Issa, S.M., Dite, T.A., O’Brien, M.T., Chen, Z.P., Galic, S.,
Signal. 5, ra42. Langendorf, C.G., Steinberg, G.R., Kemp, B.E., and Oakhill, J.S. (2014). Small
molecule drug A-769662 and AMP synergistically activate naive AMPK inde-
Rogala, K.B., Gu, X., Kedir, J.F., Abu-Remaileh, M., Bianchi, L.F., Bottino, pendent of upstream kinase signaling. Chem. Biol. 21, 619–627.
A.M.S., Dueholm, R., Niehaus, A., Overwijn, D., Fils, A.P., et al. (2019). Struc-
tural basis for the docking of mTORC1 on the lysosomal surface. Science 366, Sekiguchi, T., Kamada, Y., Furuno, N., Funakoshi, M., and Kobayashi, H.
468–475. (2014). Amino acid residues required for Gtr1p-Gtr2p complex formation and
its interactions with the Ego1p-Ego3p complex and TORC1 components in
Ross, F.A., Jensen, T.E., and Hardie, D.G. (2016a). Differential regulation by yeast. Genes Cells 19, 449–463.
AMP and ADP of AMPK complexes containing different g subunit isoforms.
Biochem. J. 473, 189–199. Settembre, C., Zoncu, R., Medina, D.L., Vetrini, F., Erdin, S., Erdin, S., Huynh,
T., Ferron, M., Karsenty, G., Vellard, M.C., et al. (2012). A lysosome-to-nucleus
Ross, F.A., MacKintosh, C., and Hardie, D.G. (2016b). AMP-activated protein signalling mechanism senses and regulates the lysosome via mTOR and
kinase: a cellular energy sensor that comes in 12 flavours. FEBS J. 283, TFEB. EMBO J. 31, 1095–1108.
2987–3001.
Shaw, R.J., Kosmatka, M., Bardeesy, N., Hurley, R.L., Witters, L.A., DePinho,
Sabatini, D.M., Erdjument-Bromage, H., Lui, M., Tempst, P., and Snyder, S.H.
(1994). RAFT1: a mammalian protein that binds to FKBP12 in a rapamycin- R.A., and Cantley, L.C. (2004). The tumor suppressor LKB1 kinase directly ac-
tivates AMP-activated kinase and regulates apoptosis in response to energy
dependent fashion and is homologous to yeast TORs. Cell 78, 35–43.
stress. Proc. Natl. Acad. Sci. USA 101, 3329–3335.
Sabers, C.J., Martin, M.M., Brunn, G.J., Williams, J.M., Dumont, F.J., Wieder-
recht, G., and Abraham, R.T. (1995). Isolation of a protein target of the Shen, K., and Sabatini, D.M. (2018). Ragulator and SLC38A9 activate the Rag
FKBP12-rapamycin complex in mammalian cells. J. Biol. Chem. 270, 815–822. GTPases through noncanonical GEF mechanisms. Proc. Natl. Acad. Sci. USA
115, 9545–9550.
Sagan, L. (1967). On the origin of mitosing cells. J. Theor. Biol. 14, 255–274.
Shen, K., Huang, R.K., Brignole, E.J., Condon, K.J., Valenstein, M.L., Chantra-
Saldivia, M., Barquilla, A., Bart, J.-M., Diaz-González, R., Hall, M.N., and Nav- nupong, L., Bomaliyamu, A., Choe, A., Hong, C., Yu, Z., and Sabatini, D.M.
arro, M. (2013). Target of rapamycin (TOR) kinase in Trypanosoma brucei: an (2018). Architecture of the human GATOR1 and GATOR1-Rag GTPases com-
extended family. Biochem. Soc. Trans. 41, 934–938. plexes. Nature 556, 64–69.

490 Cell Metabolism 31, March 3, 2020


Cell Metabolism

Review
Shen, K., Valenstein, M.L., Gu, X., and Sabatini, D.M. (2019). Arg-78 of Nprl2 in the nucleus via Ca2+/CaMKK2 signaling to enhance tumor cell survival.
catalyzes GATOR1-stimulated GTP hydrolysis by the Rag GTPases. J. Biol. Mol. Cancer Res. 16, 345–357.
Chem. 294, 2970–2975.
Vara-Ciruelos, D., Dandapani, M., Russell, F.M., Grzes, K.M., Atrih, A., Foretz,
Shertz, C.A., Bastidas, R.J., Li, W., Heitman, J., and Cardenas, M.E. (2010). M., Viollet, B., Lamont, D.J., Cantrell, D.A., and Hardie, D.G. (2019). Phenfor-
Conservation, duplication, and loss of the Tor signaling pathway in the fungal min, but not metformin, delays development of T cell acute lymphoblastic leu-
kingdom. BMC Genomics 11, 510. kemia/lymphoma via cell-autonomous AMPK activation. Cell Rep. 27, 690–
698.e694.
Shimobayashi, M., and Hall, M.N. (2014). Making new contacts: the mTOR
network in metabolism and signalling crosstalk. Nat. Rev. Mol. Cell Biol. 15, Vincent, O., Townley, R., Kuchin, S., and Carlson, M. (2001). Subcellular local-
155–162. ization of the Snf1 kinase is regulated by specific beta subunits and a novel
glucose signaling mechanism. Genes Dev. 15, 1104–1114.
Shin, H.J., Kim, H., Oh, S., Lee, J.G., Kee, M., Ko, H.J., Kweon, M.N., Won,
K.J., and Baek, S.H. (2016). AMPK-SKP2-CARM1 signalling cascade in tran- Wang, L., Lawrence, J.C., Jr., Sturgill, T.W., and Harris, T.E. (2009). Mamma-
scriptional regulation of autophagy. Nature 534, 553–557. lian target of rapamycin complex 1 (mTORC1) activity is associated with phos-
phorylation of raptor by mTOR. J. Biol. Chem. 284, 14693–14697.
Smith, F.C., Davies, S.P., Wilson, W.A., Carling, D., and Hardie, D.G. (1999).
The SNF1 kinase complex from Saccharomyces cerevisiae phosphorylates Wang, B., Zhao, L., Fish, M., Logan, C.Y., and Nusse, R. (2015a). Self-renew-
the transcriptional repressor protein Mig1p in vitro at four sites within or near ing diploid Axin2(+) cells fuel homeostatic renewal of the liver. Nature 524,
regulatory domain 1. FEBS Lett. 453, 219–223. 180–185.

Son, S.M., Park, S.J., Lee, H., Siddiqi, F., Lee, J.E., Menzies, F.M., and Ru- Wang, S., Tsun, Z.-Y.Z.Y., Wolfson, R.L., Shen, K., Wyant, G.A., Plovanich,
binsztein, D.C. (2019). Leucine signals to mTORC1 via Its metabolite acetyl- M.E., Yuan, E.D., Jones, T.D., Chantranupong, L., Comb, W., et al. (2015b).
Coenzyme A. Cell Metab. 29, 192–201.e197. Metabolism. Lysosomal amino acid transporter SLC38A9 signals arginine suf-
ficiency to mTORC1. Science 347, 188–194.
Stahmann, N., Woods, A., Carling, D., and Heller, R. (2006). Thrombin acti-
vates AMP-activated protein kinase in endothelial cells via a pathway involving Wilson, W.A., Hawley, S.A., and Hardie, D.G. (1996). Glucose repression/dere-
Ca2+/calmodulin-dependent protein kinase kinase beta. Mol. Cell. Biol. 26, pression in budding yeast: SNF1 protein kinase is activated by phosphoryla-
5933–5945. tion under derepressing conditions, and this correlates with a high AMP:ATP
ratio. Curr. Biol. 6, 1426–1434.
Stahmann, N., Woods, A., Spengler, K., Heslegrave, A., Bauer, R., Krause, S.,
Viollet, B., Carling, D., and Heller, R. (2010). Activation of AMP-activated pro- Winder, W.W., and Hardie, D.G. (1996). Inactivation of acetyl-CoA carboxylase
tein kinase by vascular endothelial growth factor mediates endothelial angio- and activation of AMP-activated protein kinase in muscle during exercise. Am.
genesis independently of nitric-oxide synthase. J. Biol. Chem. 285, J. Physiol. 270, E299–E304.
10638–10652.
Winder, W.W., Holmes, B.F., Rubink, D.S., Jensen, E.B., Chen, M., and Hol-
Stuttfeld, E., Aylett, C.H., Imseng, S., Boehringer, D., Scaiola, A., Sauer, E., loszy, J.O. (2000). Activation of AMP-activated protein kinase increases mito-
Hall, M.N., Maier, T., and Ban, N. (2018). Architecture of the human mTORC2 chondrial enzymes in skeletal muscle. J. Appl. Physiol. 88, 2219–2226.
core complex. Elife 7, e33101.
Wolfson, R.L., and Sabatini, D.M. (2017). The Dawn of the Age of Amino Acid
Sugden, C., Crawford, R.M., Halford, N.G., and Hardie, D.G. (1999a). Regula- Sensors for the mTORC1 Pathway. Cell Metab. 26, 301–309.
tion of spinach SNF1-related (SnRK1) kinases by protein kinases and phos-
phatases is associated with phosphorylation of the T loop and is regulated Wolfson, R.L., Chantranupong, L., Saxton, R.A., Shen, K., Scaria, S.M.,
by 50 -AMP. Plant J. 19, 433–439. Cantor, J.R., and Sabatini, D.M. (2016). Sestrin2 is a leucine sensor for the
mTORC1 pathway. Science 351, 43–48.
Sugden, C., Donaghy, P.G., Halford, N.G., and Hardie, D.G. (1999b). Two
SNF1-related protein kinases from spinach leaf phosphorylate and inactivate Wolfson, R.L., Chantranupong, L., Wyant, G.A., Gu, X., Orozco, J.M., Shen, K.,
3-hydroxy-3-methylglutaryl-coenzyme A reductase, nitrate reductase, and su- Condon, K.J., Petri, S., Kedir, J., Scaria, S.M., et al. (2017). KICSTOR recruits
crose phosphate synthase in vitro. Plant Physiol. 120, 257–274. GATOR1 to the lysosome and is necessary for nutrients to regulate mTORC1.
Nature 543, 438–442.
Suzuki, T., Bridges, D., Nakada, D., Skiniotis, G., Morrison, S.J., Lin, J.D., Salt-
iel, A.R., and Inoki, K. (2013). Inhibition of AMPK catabolic action by GSK3. Woods, A., Munday, M.R., Scott, J., Yang, X., Carlson, M., and Carling, D.
Mol. Cell 50, 407–419. (1994). Yeast SNF1 is functionally related to mammalian AMP-activated pro-
tein kinase and regulates acetyl-CoA carboxylase in vivo. J. Biol. Chem.
Tatebe, H., and Shiozaki, K. (2017). Evolutionary Conservation of the Compo- 269, 19509–19515.
nents in the TOR Signaling Pathways. Biomolecules 7, 77.
Woods, A., Johnstone, S.R., Dickerson, K., Leiper, F.C., Fryer, L.G., Neumann,
Thelander, M., Olsson, T., and Ronne, H. (2004). Snf1-related protein kinase 1 D., Schlattner, U., Wallimann, T., Carlson, M., and Carling, D. (2003). LKB1 is
is needed for growth in a normal day-night light cycle. EMBO J. 23, 1900–1910. the upstream kinase in the AMP-activated protein kinase cascade. Curr. Biol.
13, 2004–2008.
Thornton, C., Sardini, A., and Carling, D. (2008). Muscarinic receptor activation
of AMP-activated protein kinase inhibits orexigenic neuropeptide mRNA Woods, A., Dickerson, K., Heath, R., Hong, S.P., Momcilovic, M., Johnstone,
expression. J. Biol. Chem. 283, 17116–17122. S.R., Carlson, M., and Carling, D. (2005). Ca2+/calmodulin-dependent protein
kinase kinase-beta acts upstream of AMP-activated protein kinase in mamma-
Toyama, E.Q., Herzig, S., Courchet, J., Lewis, T.L., Jr., Losón, O.C., Hellberg, lian cells. Cell Metab. 2, 21–33.
K., Young, N.P., Chen, H., Polleux, F., Chan, D.C., and Shaw, R.J. (2016).
Metabolism. AMP-activated protein kinase mediates mitochondrial fission in Wu, N., Zheng, B., Shaywitz, A., Dagon, Y., Tower, C., Bellinger, G., Shen,
response to energy stress. Science 351, 275–281. C.H., Wen, J., Asara, J., McGraw, T.E., et al. (2013). AMPK-dependent degra-
dation of TXNIP upon energy stress leads to enhanced glucose uptake via
Treitel, M.A., Kuchin, S., and Carlson, M. (1998). Snf1 protein kinase regulates GLUT1. Mol. Cell 49, 1167–1175.
phosphorylation of the Mig1 repressor in Saccharomyces cerevisiae. Mol. Cell.
Biol. 18, 6273–6280. Wyant, G.A., Abu-Remaileh, M., Wolfson, R.L., Chen, W.W., Freinkman, E.,
Danai, L.V., Vander Heiden, M.G., and Sabatini, D.M. (2017). mTORC1 Acti-
Tsun, Z.-Y., Bar-Peled, L., Chantranupong, L., Zoncu, R., Wang, T., Kim, C., vator SLC38A9 Is Required to Efflux Essential Amino Acids from Lysosomes
Spooner, E., and Sabatini, D.M. (2013). The folliculin tumor suppressor is a and Use Protein as a Nutrient. Cell 171, 642–654.e12.
GAP for the RagC/D GTPases that signal amino acid levels to mTORC1.
Mol. Cell 52, 495–505. Xia, J., Wang, R., Zhang, T., and Ding, J. (2016). Structural insight into the argi-
nine-binding specificity of CASTOR1 in amino acid-dependent mTORC1
van Dam, T.J.P., Zwartkruis, F.J.T., Bos, J.L., and Snel, B. (2011). Evolution of signaling. Cell Discov. 2, 16035.
the TOR pathway. J. Mol. Evol. 73, 209–220.
Xiao, B., Sanders, M.J., Underwood, E., Heath, R., Mayer, F.V., Carmena, D.,
Vara-Ciruelos, D., Dandapani, M., Gray, A., Egbani, E.O., Evans, A.M., and Jing, C., Walker, P.A., Eccleston, J.F., Haire, L.F., et al. (2011). Structure of
Hardie, D.G. (2018). Genotoxic damage activates the AMPK-alpha1 isoform mammalian AMPK and its regulation by ADP. Nature 472, 230–233.

Cell Metabolism 31, March 3, 2020 491


Cell Metabolism

Review
Xiao, B., Sanders, M.J., Carmena, D., Bright, N.J., Haire, L.F., Underwood, E., Yuan, H.-X., Wang, Z., Yu, F.-X., Li, F., Russell, R.C., Jewell, J.L., and Guan,
Patel, B.R., Heath, R.B., Walker, P.A., Hallen, S., et al. (2013). Structural basis K.-L. (2015). NLK phosphorylates Raptor to mediate stress-induced mTORC1
of AMPK regulation by small molecule activators. Nat. Commun. 4, 3017. inhibition. Genes Dev. 29, 2362–2376.

Xin, F.J., Wang, J., Zhao, R.Q., Wang, Z.X., and Wu, J.W. (2013). Coordinated Zeng, L., Fagotto, F., Zhang, T., Hsu, W., Vasicek, T.J., Perry, W.L., 3rd, Lee,
regulation of AMPK activity by multiple elements in the a-subunit. Cell Res. 23, J.J., Tilghman, S.M., Gumbiner, B.M., and Costantini, F. (1997). The mouse
1237–1240. Fused locus encodes Axin, an inhibitor of the Wnt signaling pathway that reg-
ulates embryonic axis formation. Cell 90, 181–192.
Xue, R.Q., Sun, L., Yu, X.J., Li, D.L., and Zang, W.J. (2016). Vagal nerve stim-
ulation improves mitochondrial dynamics via an M3 receptor/CaMKKbeta/ Zeqiraj, E., Filippi, B.M., Deak, M., Alessi, D.R., and van Aalten, D.M. (2009).
AMPK pathway in isoproterenol-induced myocardial ischaemia. J. Cell. Mol. Structure of the LKB1-STRAD-MO25 complex reveals an allosteric mecha-
Med. 21, 58–71. nism of kinase activation. Science 326, 1707–1711.

Zhang, T., Péli-Gulli, M.-P., Yang, H., De Virgilio, C., and Ding, J. (2012). Ego3
Yan, Y., Zhou, X.E., Novick, S.J., Shaw, S.J., Li, Y., Brunzelle, J.S., Hitoshi, Y., functions as a homodimer to mediate the interaction between Gtr1-Gtr2 and
Griffin, P.R., Xu, H.E., and Melcher, K. (2019). Structures of AMP-activated Ego1 in the ego complex to activate TORC1. Structure 20, 2151–2160.
protein kinase bound to novel pharmacological activators in phosphorylated,
non-phosphorylated, and nucleotide-free states. J. Biol. Chem. 294, 953–967. Zhang, Y.L., Guo, H., Zhang, C.S., Lin, S.Y., Yin, Z., Peng, Y., Luo, H., Shi, Y.,
Lian, G., Zhang, C., et al. (2013). AMP as a low-energy charge signal autono-
Yang, Y., Atasoy, D., Su, H.H., and Sternson, S.M. (2011). Hunger states mously initiates assembly of AXIN-AMPK-LKB1 complex for AMPK activation.
switch a flip-flop memory circuit via a synaptic AMPK-dependent positive Cell Metab. 18, 546–555.
feedback loop. Cell 146, 992–1003.
Zhang, C.S., Jiang, B., Li, M., Zhu, M., Peng, Y., Zhang, Y.L., Wu, Y.Q., Li, T.Y.,
Yang, H., Rudge, D.G., Koos, J.D., Vaidialingam, B., Yang, H.J., and Pavletich, Liang, Y., Lu, Z., et al. (2014). The lysosomal v-ATPase-Ragulator complex is a
N.P. (2013). mTOR kinase structure, mechanism and regulation. Nature 497, common activator for AMPK and mTORC1, acting as a switch between catab-
217–223. olism and anabolism. Cell Metab. 20, 526–540.

Yang, H., Wang, J., Liu, M., Chen, X., Huang, M., Tan, D., Dong, M.-Q., Wong, Zhang, C.S., Hawley, S.A., Zong, Y., Li, M., Wang, Z., Gray, A., Ma, T., Cui, J.,
C.C.L., Wang, J., Xu, Y., and Wang, H.W. (2016). 4.4 Å Resolution Cryo-EM Feng, J.W., Zhu, M., et al. (2017). Fructose-1,6-bisphosphate and aldolase
structure of human mTOR Complex 1. Protein Cell 7, 878–887. mediate glucose sensing by AMPK. Nature 548, 112–116.

Yang, H., Jiang, X., Li, B., Yang, H.J., Miller, M., Yang, A., Dhar, A., and Pav- Zhao, R.Q. (2019). Expression, purification and characterization of the plant
letich, N.P. (2017). Mechanisms of mTORC1 activation by RHEB and inhibition Snf1-related protein kinase 1 from Escherichia coli. Protein Expr. Purif.
by PRAS40. Nature 552, 368–373. 162, 24–31.

Zoncu, R., Bar-Peled, L., Efeyan, A., Wang, S., Sancak, Y., and Sabatini, D.M.
Yeh, L.A., Lee, K.H., and Kim, K.H. (1980). Regulation of rat liver acetyl-CoA (2011). mTORC1 senses lysosomal amino acids through an inside-out mech-
carboxylase. Regulation of phosphorylation and inactivation of acetyl-CoA anism that requires the vacuolar H(+)-ATPase. Science 334, 678–683.
carboxylase by the adenylate energy charge. J. Biol. Chem. 255, 2308–2314.
Zong, H., Ren, J.M., Young, L.H., Pypaert, M., Mu, J., Birnbaum, M.J., and
Yoon, M.S., Sun, Y., Arauz, E., Jiang, Y., and Chen, J. (2011). Phosphatidic Shulman, G.I. (2002). AMP kinase is required for mitochondrial biogenesis in
acid activates mammalian target of rapamycin complex 1 (mTORC1) kinase skeletal muscle in response to chronic energy deprivation. Proc. Natl. Acad.
by displacing FK506 binding protein 38 (FKBP38) and exerting an allosteric ef- Sci. USA 99, 15983–15987.
fect. J. Biol. Chem. 286, 29568–29574.
Zong, Y., Zhang, C.S., Li, M., Wang, W., Wang, Z., Hawley, S.A., Ma, T., Feng,
Yoon, M.S., Son, K., Arauz, E., Han, J.M., Kim, S., and Chen, J. (2016). Leucyl- J.W., Tian, X., Qi, Q., et al. (2019). Hierarchical activation of compartmentalized
tRNA synthetase activates Vps34 in amino acid-sensing mTORC1 signaling. pools of AMPK depends on severity of nutrient or energy stress. Cell Res. 29,
Cell Rep. 16, 1510–1517. 460–473.

492 Cell Metabolism 31, March 3, 2020

You might also like